[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,783)

Search Parameters:
Keywords = MODIS data

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 326 KiB  
Review
MODY Only Monogenic? A Narrative Review of the Novel Rare and Low-Penetrant Variants
by Iderina Hasballa and Davide Maggi
Int. J. Mol. Sci. 2024, 25(16), 8790; https://doi.org/10.3390/ijms25168790 (registering DOI) - 13 Aug 2024
Viewed by 102
Abstract
Maturity-onset diabetes of the young (MODY) represents the most frequent form of monogenic diabetes mellitus (DM), currently classified in 14 distinct subtypes according to single gene mutations involved in the differentiation and function of pancreatic β-cells. A significant proportion of MODY has unknown [...] Read more.
Maturity-onset diabetes of the young (MODY) represents the most frequent form of monogenic diabetes mellitus (DM), currently classified in 14 distinct subtypes according to single gene mutations involved in the differentiation and function of pancreatic β-cells. A significant proportion of MODY has unknown etiology, suggesting that the genetic landscape is still to be explored. Recently, novel potentially MODY-causal genes, involved in the differentiation and function of β-cells, have been identified, such as RFX6, NKX2.2, NKX6.1, WFS1, PCBD1, MTOR, TBC1D4, CACNA1E, MNX1, AKT2, NEUROG3, EIF2AK3, GLIS3, HADH, and PTF1A. Genetic and clinical features of MODY variants remain highly heterogeneous, with no direct genotype–phenotype correlation, especially in the low-penetrant subtypes. This is a narrative review of the literature aimed at describing the current state-of-the-art of the novel likely MODY-associated variants. For a deeper understanding of MODY complexity, we also report some related controversies concerning the etiological role of some of the well-known pathological genes and MODY inheritance pattern, as well as the rare association of MODY with autoimmune diabetes. Due to the limited data available, the assessment of MODY-related genes pathogenicity remains challenging, especially in the setting of rare and low-penetrant subtypes. In consideration of the crucial importance of an accurate diagnosis, prognosis and management of MODY, more studies are warranted to further investigate its genetic landscape and the genotype–phenotype correlation, as well as the pathogenetic contribution of the nongenetic modifiers in this cohort of patients. Full article
(This article belongs to the Special Issue Molecular Research on Diabetes)
22 pages, 11626 KiB  
Article
Assessment of Semi-Automated Techniques for Crop Mapping in Chile Based on Global Land Cover Satellite Data
by Matías Volke, María Pedreros-Guarda, Karen Escalona, Eduardo Acuña and Raúl Orrego
Remote Sens. 2024, 16(16), 2964; https://doi.org/10.3390/rs16162964 (registering DOI) - 12 Aug 2024
Viewed by 242
Abstract
In recent years, the Chilean agricultural sector has undergone significant changes, but there is a lack of data that can be used to accurately identify these transformations. A study was conducted to assess the effectiveness of different spatial resolutions used by global land [...] Read more.
In recent years, the Chilean agricultural sector has undergone significant changes, but there is a lack of data that can be used to accurately identify these transformations. A study was conducted to assess the effectiveness of different spatial resolutions used by global land cover products (MODIS, ESA and Dynamic World (DW)), in addition to the demi-automated methods applied to them, for the identification of agricultural areas, using the publicly available agricultural survey for 2021. It was found that lower-spatial-resolution collections consistently underestimated crop areas, while collections with higher spatial resolutions overestimated them. The low-spatial-resolution collection, MODIS, underestimated cropland by 46% in 2021, while moderate-resolution collections, such as ESA and DW, overestimated cropland by 39.1% and 93.8%, respectively. Overall, edge-pixel-filtering and a machine learning semi-automated reclassification methodology improved the accuracy of the original global collections, with differences of only 11% when using the DW collection. While there are limitations in certain regions, the use of global land cover collections and filtering methods as training samples can be valuable in areas where high-resolution data are lacking. Future research should focus on validating and adapting these approaches to ensure their effectiveness in sustainable agriculture and ecosystem conservation on a global scale. Full article
(This article belongs to the Special Issue Application of Satellite and UAV Data in Precision Agriculture)
Show Figures

Figure 1

Figure 1
<p>Study area. On the left, Chile with the Ñuble region is marked in black. On the right, a zoomed image of the Ñuble region and the names of its communes are shown.</p>
Full article ">Figure 2
<p>Flow diagram representing the methodology of this work.</p>
Full article ">Figure 3
<p>Agricultural land area in km<sup>2</sup>, (<b>a</b>) per commune in the Ñuble region and (<b>b</b>) the total in the region. Year: 2021. The calculations were retrieved from the following databases: an agricultural survey (AS), MODIS, ESA, Dynamic World (DW) and improved versions of the latter three (v2).</p>
Full article ">Figure 4
<p>Agricultural area in km<sup>2</sup> per commune in the Ñuble region. (<b>a</b>) Agricultural survey and (<b>b</b>) original ESA dataset, as the most precise original dataset from the three tested. (<b>c</b>) Filtered and reclassified DW, that is, DW version 2 (v2), as the most accurate filtered and reclassified dataset.</p>
Full article ">Figure 4 Cont.
<p>Agricultural area in km<sup>2</sup> per commune in the Ñuble region. (<b>a</b>) Agricultural survey and (<b>b</b>) original ESA dataset, as the most precise original dataset from the three tested. (<b>c</b>) Filtered and reclassified DW, that is, DW version 2 (v2), as the most accurate filtered and reclassified dataset.</p>
Full article ">Figure 5
<p>Zoomed image of the pixel reduction through different stages of quality filtering of <a href="#remotesensing-16-02964-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure A1
<p>Maps of agricultural mask retrieved from MODIS dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A2
<p>Maps of agricultural mask retrieved from DW dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A3
<p>Maps of agricultural mask retrieved from ESA dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A3 Cont.
<p>Maps of agricultural mask retrieved from ESA dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A4
<p>Maps of agricultural mask retrieved from CONAF dataset for the year 2021.</p>
Full article ">
23 pages, 29093 KiB  
Article
Utilizing the Google Earth Engine for Agricultural Drought Conditions and Hazard Assessment Using Drought Indices in the Najd Region, Sultanate of Oman
by Mohammed S. Al Nadabi, Paola D’Antonio, Costanza Fiorentino, Antonio Scopa, Eltaher M. Shams and Mohamed E. Fadl
Remote Sens. 2024, 16(16), 2960; https://doi.org/10.3390/rs16162960 - 12 Aug 2024
Viewed by 224
Abstract
Accurately evaluating drought and its effects on the natural environment is difficult in regions with limited climate monitoring stations, particularly in the hyper-arid region of the Sultanate of Oman. Rising global temperatures and increasing incidences of insufficient precipitation have turned drought into a [...] Read more.
Accurately evaluating drought and its effects on the natural environment is difficult in regions with limited climate monitoring stations, particularly in the hyper-arid region of the Sultanate of Oman. Rising global temperatures and increasing incidences of insufficient precipitation have turned drought into a major natural disaster worldwide. In Oman, drought constitutes a major threat to food security. In this study, drought indices (DIs), such as temperature condition index (TCI), vegetation condition index (VCI), and vegetation health index (VHI), which integrate data on drought streamflow, were applied using moderate resolution imaging spectroradiometer (MODIS) data and the Google Earth Engine (GEE) platform to monitor agricultural drought and assess the drought risks using the drought hazard index (DHI) during the period of 2001–2023. This approach allowed us to explore the spatial and temporal complexities of drought patterns in the Najd region. As a result, the detailed analysis of the TCI values exhibited temporal variations over the study period, with notable minimum values observed in specific years (2001, 2005, 2009, 2010, 2014, 2015, 2016, 2017, 2019, 2020, and 2021), and there was a discernible trend of increasing temperatures from 2014 to 2023 compared to earlier years. According to the VCI index, several years, including 2001, 2003, 2006, 2008, 2009, 2013, 2015, 2016, 2017, 2018, 2020, 2021, 2022, and 2023, were characterized by mild drought conditions. Except for 2005 and 2007, all studied years were classified as moderate drought years based on the VHI index. The Pearson correlation coefficient analysis (PCA) was utilized to observe the correlation between DIs, and a high positive correlation between VHI and VCI (0.829, p < 0.01) was found. Based on DHI index spatial analysis, the northern regions of the study area faced the most severe drought hazards, with severity gradually diminishing towards the south and east, and approximately 44% of the total area fell under moderate drought risk, while the remaining 56% was classified as facing very severe drought risk. This study emphasizes the importance of continued monitoring, proactive measures, and effective adaptation strategies to address the heightened risk of drought and its impacts on local ecosystems and communities. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area (The Najd region, Sultanate of Oman).</p>
Full article ">Figure 2
<p>Schematic overview of Google Earth Engine (GEE) data processing.</p>
Full article ">Figure 3
<p>The methodological framework used in this study.</p>
Full article ">Figure 4
<p>TCI, time series plot of the Najd region derived using GEE and MODIS images.</p>
Full article ">Figure 5
<p>LST variation trends during the 2001–2023 period at (<b>a</b>) Marmul and (<b>b</b>) Thumrait meteorological stations.</p>
Full article ">Figure 5 Cont.
<p>LST variation trends during the 2001–2023 period at (<b>a</b>) Marmul and (<b>b</b>) Thumrait meteorological stations.</p>
Full article ">Figure 6
<p>VCI, time series plot of the Najd region derived using GEE and MODIS images.</p>
Full article ">Figure 7
<p>VHI, time series plot of the Najd region derived using GEE and MODIS images.</p>
Full article ">Figure 8
<p>Descriptive statistics of VCI, TCI, and VHI values during the 2001–2023 period at the Najd region.</p>
Full article ">Figure 9
<p>Spatial distribution of TCI, VCI, and VHI in the Najd region retrieved from MODIS Satellites for the period of 2001 (<b>a</b>)–2023 (<b>w</b>).</p>
Full article ">Figure 9 Cont.
<p>Spatial distribution of TCI, VCI, and VHI in the Najd region retrieved from MODIS Satellites for the period of 2001 (<b>a</b>)–2023 (<b>w</b>).</p>
Full article ">Figure 9 Cont.
<p>Spatial distribution of TCI, VCI, and VHI in the Najd region retrieved from MODIS Satellites for the period of 2001 (<b>a</b>)–2023 (<b>w</b>).</p>
Full article ">Figure 9 Cont.
<p>Spatial distribution of TCI, VCI, and VHI in the Najd region retrieved from MODIS Satellites for the period of 2001 (<b>a</b>)–2023 (<b>w</b>).</p>
Full article ">Figure 9 Cont.
<p>Spatial distribution of TCI, VCI, and VHI in the Najd region retrieved from MODIS Satellites for the period of 2001 (<b>a</b>)–2023 (<b>w</b>).</p>
Full article ">Figure 10
<p>The spatial distribution of drought hazards in the Najd region over the time period.</p>
Full article ">
14 pages, 7934 KiB  
Article
Assimilating Satellite-Derived Snow Cover and Albedo Data to Improve 3-D Weather and Photochemical Models
by Colleen Jones, Huy Tran, Trang Tran and Seth Lyman
Atmosphere 2024, 15(8), 954; https://doi.org/10.3390/atmos15080954 (registering DOI) - 10 Aug 2024
Viewed by 217
Abstract
During wintertime temperature inversion episodes, ozone in the Uinta Basin sometimes exceeds the standard of 70 ppb set by the US Environmental Protection Agency. Since ozone formation depends on sunlight, and less sunlight is available during winter, wintertime ozone can only form if [...] Read more.
During wintertime temperature inversion episodes, ozone in the Uinta Basin sometimes exceeds the standard of 70 ppb set by the US Environmental Protection Agency. Since ozone formation depends on sunlight, and less sunlight is available during winter, wintertime ozone can only form if snow cover and albedo are high. Researchers have encountered difficulties replicating high albedo values in 3-D weather and photochemical transport model simulations for winter episodes. In this study, a process to assimilate MODIS satellite data into WRF and CAMx models was developed, streamlined, and tested to demonstrate the impacts of data assimilation on the models’ performance. Improvements to the WRF simulation of surface albedo and snow cover were substantial. However, the impact of MODIS data assimilation on WRF performance for other meteorological quantities was minimal, and it had little impact on ozone concentrations in the CAMx photochemical transport model. The contrast between the data assimilation and reference cases was greater for a period with no new snow since albedo appears to decrease too rapidly in default WRF and CAMx configurations. Overall, the improvement from MODIS data assimilation had an observed enhancement in the spatial distribution and temporal evolution of surface characteristics on meteorological quantities and ozone production. Full article
Show Figures

Figure 1

Figure 1
<p>WRF one-way nested 12-4-1.33 km domains (<b>A</b>) and details of a 1.33 km domain, including topography and location of oil and gas wells (<b>B</b>). The white rectangle is Domain 2 and the red rectangle is Domain 3 from Table 4.</p>
Full article ">Figure 2
<p>Diagram of the MODIS data assimilation into the WRF and CAMx models.</p>
Full article ">Figure 3
<p>Comparison of the surface albedo fraction obtained in simulations using the WRF default configuration (<b>left</b>) and MODIS data assimilation (<b>right</b>). (Thin black lines = county outlines; heavy black outline = Uinta Basin ozone nonattainment area; and two-letter abbreviations = air quality monitoring stations.)</p>
Full article ">Figure 4
<p>Comparison of snow cover fraction (SNOWC) obtained in simulations using the WRF default configuration (<b>left</b>) and MODIS data assimilation (<b>right</b>). (Thin black lines = county outlines; heavy black outline = Uinta Basin ozone nonattainment area; and two-letter abbreviations = air quality monitoring stations.)</p>
Full article ">Figure 5
<p>Comparison of snow cover fraction, snow water equivalent, and snow depth using the WRF default configuration (REF) and MODIS data assimilation (MODIS). Green bars show periods where WRF reinitialized snow characteristics using the SNOWDAS dataset.</p>
Full article ">Figure 6
<p>Comparison of planetary boundary layer height (P.B.L.H.) and lapse rate using the WRF default configuration (REF) and MODIS data assimilation (MODIS).</p>
Full article ">Figure 7
<p>Comparison of photolysis rates simulated by CAMx using the default configuration (REF) and MODIS data assimilation (MODIS). (Green line = a new snow event.)</p>
Full article ">Figure 8
<p>Comparison of ozone at Ouray as simulated by CAMx using the default configuration (REF) and MODIS data assimilation (MODIS). (Red dash line = EPA National Ambient Air Quality Standard (NAAQS) for ozone).</p>
Full article ">
19 pages, 9008 KiB  
Article
The Carpathian Agriculture in Poland in Relation to Other EU Countries, Ukraine and the Environmental Goals of the EU CAP 2023–2027
by Marek Zieliński, Artur Łopatka, Piotr Koza and Barbara Gołębiewska
Agriculture 2024, 14(8), 1325; https://doi.org/10.3390/agriculture14081325 - 9 Aug 2024
Viewed by 267
Abstract
This study discusses the issue of determining the direction and strength of changes taking place in the structure of agricultural land in the mountain and foothill areas of the Carpathians in Poland in comparison with Slovakia, Romania and Ukraine. The most important financial [...] Read more.
This study discusses the issue of determining the direction and strength of changes taking place in the structure of agricultural land in the mountain and foothill areas of the Carpathians in Poland in comparison with Slovakia, Romania and Ukraine. The most important financial institutional measures dedicated to the protection of the natural environment in Polish agriculture in the Areas facing Natural and other specific Constraints (ANCs) mountain and foothill in the first year of the CAP 2023–2027 were also established. Satellite data from 2001 to 2022 were used. The analyses used the land use classification MCD12Q1 provided by NASA and were made on the basis of satellite imagery collections from the MODIS sensor placed on two satellites: TERRA and AQUA. In EU countries, a decreasing trend in agricultural areas has been observed in areas below 350 m above sea level. In areas above 350 m, this trend weakened or even turned into an upward trend. Only in Ukraine was a different trend observed. It was found that in Poland, the degree of involvement of farmers from mountain and foothill areas in implementing financial institutional measures dedicated to protecting the natural environment during the study period was not satisfactory. Full article
Show Figures

Figure 1

Figure 1
<p>Scheme of the analysis of agriculture within separate groups of communes due to the fact and nuisance of ANCs mountain and foothill in Poland. Source: own study.</p>
Full article ">Figure 2
<p>Distribution of communes with different shares of ANCs mountain and foothill in Poland. Source: own study ISSPC SRI; IAFE NRI.</p>
Full article ">Figure 3
<p>Land use in the Carpathians in 2001 and 2022. Source: own study based on MODIS.</p>
Full article ">Figure 4
<p>Trends in the percentage share [%] of the total agricultural area and cropland in the total area of land in the Carpathians in 2001–2022. Source: own study based on MODIS.</p>
Full article ">Figure 5
<p>Number of farms participating in practices under eco-schemes, in organic and agri–environment–climate measures in communes with different shares of ANCs mountain and foothill in Poland in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 6
<p>Share of [%] farms with eco-schemes in total number of farms in communes with ANCs mountain and foothill in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 7
<p>Share of [%] farms with organic and agri–environmental–climate measure in total number of farms in communes with ANCs mountain and foothill in 2023.</p>
Full article ">Figure 8
<p>Agricultural area covered by practices under eco-schemes, ecological and agri–environment–climate measures in communes with different shares of ANCs mountain and foothill in Poland in 2023. Source: own study based on ARMA.</p>
Full article ">Figure 9
<p>Share [%] of UAA in farms with eco-schemes in total UAA in communes with ANCs mountain and foothill in 2023.</p>
Full article ">Figure 10
<p>Share [%] of UAA covered by organic and agri–environmental–climate measures in total UAA in communes with ANCs mountain and foothill in 2023.</p>
Full article ">
27 pages, 5829 KiB  
Article
Monitoring Changes in the Enhanced Vegetation Index to Inform the Management of Forests
by Peter S. Rodriguez, Amanda M. Schwantes, Andrew Gonzalez and Marie-Josée Fortin
Remote Sens. 2024, 16(16), 2919; https://doi.org/10.3390/rs16162919 - 9 Aug 2024
Viewed by 283
Abstract
In the absence of forest ecosystem time series data, monitoring proxies such as the enhanced vegetation index (EVI) can inform the capacity of forests to provide ecosystem services. We used MODIS-derived EVI at 250 m and 16-day resolution and Breaks for Additive and [...] Read more.
In the absence of forest ecosystem time series data, monitoring proxies such as the enhanced vegetation index (EVI) can inform the capacity of forests to provide ecosystem services. We used MODIS-derived EVI at 250 m and 16-day resolution and Breaks for Additive and Seasonal Trend (BFAST) algorithms to monitor forest EVI changes (breaks and trends) in and around the Algonquin Provincial Park (Ontario, Canada) from 2003 to 2022. We found that relatively little change occurred in forest EVI pixels and that most of the change occurred in non-protected forest areas. Only 5.3% (12,348) of forest pixels experienced one or more EVI breaks and 27.8% showed detectable EVI trends. Most breaks were negative (11,969, 75.3%; positive breaks: 3935, 24.7%) with a median magnitude of change of −755.5 (median positive magnitude: 722.6). A peak of negative breaks (2487, 21%) occurred in the year 2013 while no clear peak was seen among positive breaks. Most breaks (negative and positive) and trends occurred in the eastern region of the study area. Boosted regression trees revealed that the most important predictors of the magnitude of change were forest age, summer droughts, and warm winters. These were among the most important variables that explained the magnitude of negative (R2 = 0.639) and positive breaks (R2 = 0.352). Forest composition and protection status were only marginally important. Future work should focus on assessing spatial clusters of EVI breaks and trends to understand local drivers of forest vegetation health and their potential relation to forest ecosystem services. Full article
Show Figures

Figure 1

Figure 1
<p>The study area is the Algonquin Provincial Park and the surrounding area, which we refer to as the Algonquin Greater Park Ecosystem (AGPE). Elevation (in meters above sea level, masl) is shown in (<b>a</b>). The forest’s mean age in 2019 is shown in (<b>b</b>). Protected areas within the AGPE are shown in (<b>c</b>). About 16% of forest pixels belong to protected areas. The geographic distribution of three disturbance agents in the period 2002–2020 is shown in (<b>d</b>). The gray color represents forested pixels and white non-forest pixels. The dashed black line shows the study area footprint (≈15,000 km<sup>2</sup> or 1.5 M ha). The solid black line shows Algonquin Provincial Park’s footprint. The two perpendicular dashed lines in (<b>a</b>) divide the study area into quadrants (northwest, NW; northeast, NE; southwest, SW; and southeast, SE) to ease the description of the results. Map projection is NAD83 Statistics Canada Lambert, EPSG: 3347.</p>
Full article ">Figure 2
<p>Main methodological steps used in this study. After downloading MODIS EVI data, pixels were filtered to only keep pixels with good quality data, within forested areas. EVI time series were then created and processed with three BFAST algorithms: bfast, bfast01, and bfastclassify. The bfast algorithm decomposes a time series into a seasonal component, trend, and noise. Using piece-wise linear regression on the trend component, it detects one or more breaks (if any). Here, we use three main outputs provided by bfast: type of break, magnitude of break, and time of break with 95% confidence intervals (CIs) The bfast01 algorithm runs a seasonally adjusted regression model on the ts and only detects the major break (if any). The bfastclassify algorithm then uses bfast01’s output to classify trends into one of eight possible trend types (<a href="#remotesensing-16-02919-f0A1" class="html-fig">Figure A1</a>). Only the magnitude of break values estimated by bfast was used in boosted regression trees (XGBoost models) to explore their relationship with predictor variables (dashed box at bottom, Equation (<a href="#FD1-remotesensing-16-02919" class="html-disp-formula">1</a>) in main text). Satellite icon from <a href="http://flaticon.com" target="_blank">flaticon.com</a> (accessed on 29 April 2022).</p>
Full article ">Figure 3
<p>Spatial distribution of EVI negative and positive breaks in the AGPE from 2003 to 2022. Most of the breaks were found in the eastern half of the AGPE and more particularly in the northeast quadrant. There were 11,871 pixels with negative breaks (red pixels) and 3893 pixels with positive breaks (cyan pixels). These breaks were estimated with the bfast algorithm. The dashed black line shows the study area footprint. The solid black line shows Algonquin Provincial Park’s footprint. The two perpendicular dashed lines divide the study area into four quadrants (northwest, NW; northeast, NE; southwest, SW; and southeast, SE) to ease the interpretation of results. Map projection is NAD83 Statistics Canada Lambert, EPSG: 3347.</p>
Full article ">Figure 4
<p>Spatial distribution of EVI trend types in the AGPE from 2003 to 2022. These trends were produced with the bfastclassify algorithm. The trend types are those proposed by de Jong et al. [<a href="#B57-remotesensing-16-02919" class="html-bibr">57</a>]. Abbreviations—MIG: monotonic increasing, greening trend (<span class="html-italic">n</span> = 33,683); MDB: monotonic decreasing, browning trend (<span class="html-italic">n</span> = 4981); IInb: interruption, increasing trend with a negative break (<span class="html-italic">n</span> = 11,637); RBG: reversal, browning to greening trend (<span class="html-italic">n</span> = 11,654). (Trends MIGpb, MDBnb, IDpb, and RGB are not shown given their low percentages). The dashed black line shows the study area footprint. The solid black line shows Algonquin Provincial Park’s footprint. The two perpendicular dashed lines divide the study area into four quadrants (northwest, NW; northeast, NE; southwest, SW; and southeast, SE) to ease the interpretation of results. Map projection across all figures is NAD83 Statistics Canada Lambert, EPSG: 3347.</p>
Full article ">Figure 5
<p>Predictors of the magnitude of EVI breaks (negative breaks in red and positive breaks in cyan tone boxes) from 2003 to 2022. The ranking reflects feature importance using the gain metric as estimated by XGBoost models. The same five predictors are in the top five but with slightly different rankings (connecting lines with slopes) except for the summer climate moisture index with a 3-year lag, which ranks fourth in both (connecting line with no slope). Forest protection status is low-ranking for both types of breaks. The XGBoost models were run with subsets of the detected breaks (negative breaks, <span class="html-italic">n</span> = 116 records; positive breaks, <span class="html-italic">n</span> = 3263 records).</p>
Full article ">Figure A1
<p>Schematic of break and trend types employed in this study. A time series (ts) can be characterized by the presence or absence of breaks and trends. Breaks represent abrupt changes in a ts whereas trends represent gradual changes. Here, we refer to three major groups of trends: monotonic (<b>a</b>,<b>b</b>), interruption (<b>c</b>), and reversal trends (<b>d</b>). Similarly, we refer to two break types: negative (red downward arrow) and positive (green upward arrow) breaks. Trends can be monotonic, either increasing (green slopes) or decreasing (orange slopes). The former are referred to as greening trends and the latter as browning trends. Monotonic trends can show no breaks (<b>a</b>) or show one or more breaks (negative or positive) but in concordance with the slope of the trend segments (<b>b</b>). Conversely, interruption (<b>c</b>) and reversal (<b>d</b>) trends are characterized by having a break type in discordance with the slope of the trend segments. Interruption trends can have two positive trend segments divided by a negative break and vice versa (two negative trend segments divided by a positive break). Reversal trends have opposite trend segments divided by a negative or positive break. Lastly, some ts may not change or show changes that are too small to be detected with the methods employed (horizontal gray dashed line in (<b>a</b>)). Here, we use the trend classification proposed by de Jong et al. [<a href="#B57-remotesensing-16-02919" class="html-bibr">57</a>]—MIG: monotonic increasing, greening trend (without breaks) (bottom line in (<b>a</b>)); MDB: monotonic decreasing, browning trend (without breaks) (top line in (<b>a</b>)); MIGpb: monotonic increasing, greening trend with a positive break (top set of lines in (<b>b</b>)); MDBnb: monotonic decreasing, browning trend with a negative break (bottom set of lines in (<b>b</b>)); IInb: interruption, the increasing trend with a negative break (top set of lines in (<b>c</b>)); IDpb: interruption, decreasing trend with a positive break (bottom set of lines in (<b>c</b>)); RGB: reversal, greening to browning trend (top two sets of lines in (<b>d</b>)); RBG: reversal, browning to greening trend (bottom two sets of lines in (<b>d</b>)).</p>
Full article ">Figure A2
<p>Density plots of forest EVI magnitude of breaks in the AGPE from 2003 to 2022. The number of breaks and their magnitudes broken down by year are shown. Extreme magnitude values have been omitted to aid visualization. Vertical lines show medians—solid: yearly; dashed: entire time series. The total number of breaks was 15,904 (11,969 negative and 3935 positive). The time of break was rounded up prior to plotting which caused 2003 breaks (8 negative and 18 positive) to be part of 2004. No breaks were detected in 2022.</p>
Full article ">Figure A3
<p>Ranking of all predictors (features) used in the XGBoost models. Panel (<b>a</b>) shows predictors of negative break magnitudes and panel (<b>b</b>) shows those of positive break magnitudes. These models were run with subsets of the detected breaks (negative breaks, <span class="html-italic">n</span> = 116 records; positive, <span class="html-italic">n</span> = 3263 records). Variable abbreviations—for_age: forest age; dd5_wt: winter degree days above 5 °C; cmi_sm: summer climate moisture index; for_con: percentage of conifers; for_pro_0: non-protected forest; lag#: 1-, 2- or 3-year lags. The protected forest variable is not present in (<b>a</b>) given its lack of importance in explaining negative breaks.</p>
Full article ">Figure A4
<p>Partial dependence plots of magnitude of negative break predictors (features). The relationships between predictors and response (magnitude of EVI breaks) variables were mostly non-linear. Plots were created from the output of XGBoost. The model was run with a subset of all detected negative breaks (<span class="html-italic">n</span> = 116). The values on the y-axes are absolute values of negative magnitudes. Variable abbreviations—for_age: forest age; dd5_wt: winter degree days above 5 °C; cmi_sm: summer climate moisture index; for_con: percentage of conifers; for_pro_0: non-protected forest equals 1, protected equals 0; lag#: 1-, 2- or 3-year lags.</p>
Full article ">Figure A5
<p>Partial dependence plots of magnitude of positive break predictors (features). The relationships between predictors and response (magnitude of EVI breaks) variables were mostly non-linear. Plots were created from the output of XGBoost. The model was run with a subset of all the detected positive breaks (<span class="html-italic">n</span> = 3263). Variable abbreviations—for_age: forest age; dd5_wt: winter degree days above 5 °C; cmi_sm: summer climate moisture index; for_con: percentage of conifers; for_pro_0: non-protected forest equals 1, protected equals 0; lag#: 1-, 2- or 3-year lags.</p>
Full article ">Figure A6
<p>Geographic distribution of trends in Algonquin Park and the surrounding area, which we refer to as the Algonquin Greater Park Ecosystem (AGPE). All maps show trends that were derived from the output of bfastclassify. Compared to greening trends (MIG) which occurred throughout the AGPE (<b>a</b>), browning trends (MDB) mostly occurred in the NE quadrant (<b>b</b>). Most increasing trends with negative breaks (interruptions, IInb) occurred in the NW quadrant (<b>c</b>) while most of the relatively few decreasing trends with positive breaks (interruptions, IDpb) occurred in the NE quadrant (<b>c</b>). Notably, browning to greening reverse trends (RBG) co-occurred with browning trends in the NE quadrant (<b>d</b>). The dashed black line shows the study area footprint (1.5 M ha). The solid black line shows Algonquin Provincial Park’s footprint. The perpendicular dashed lines divide the study area into four quadrants (northwest, NW; northeast, NE; southwest, SW; and southeast, SE) to ease the description of the results. Map projection across all figures is NAD83 Statistics Canada Lambert, EPSG: 3347.</p>
Full article ">
20 pages, 10187 KiB  
Article
Finding Oasis Cold Island Footprints Based on a Logistic Model—A Case Study in the Ejina Oasis
by Wentong Wu and Rensheng Chen
Remote Sens. 2024, 16(16), 2895; https://doi.org/10.3390/rs16162895 - 8 Aug 2024
Viewed by 219
Abstract
Oases play a crucial role in arid regions within the human–environmental system, holding significant ecological and biological importance. The Oasis Cold Island Effect (OCIE) represents a distinct climatic feature of oases and serves as a vital metric for assessing oasis ecosystems. Previous studies [...] Read more.
Oases play a crucial role in arid regions within the human–environmental system, holding significant ecological and biological importance. The Oasis Cold Island Effect (OCIE) represents a distinct climatic feature of oases and serves as a vital metric for assessing oasis ecosystems. Previous studies have overlooked the spatial extent of the Oasis Cold Island Effect (OCIE), specifically the boundary delineating areas influenced and unaffected by oases. This boundary is defined as the Oasis Cold Island Footprint (OCI FP). Utilizing Logistic modeling and MODIS data products, OCI FPs were calculated for the Ejina Oasis from 2000 to 2019. The assessment results underscore the accuracy and feasibility of the methodology, indicating its potential applicability to other oases. Spatial and temporal distributions of OCI FPs and the intensity of the Oasis Cold Island Effect Intensity (OCIEI) in the Ejina Oasis were analyzed, yielding the following findings: (1) OCI FP area and complexity were smallest in summer and largest in autumn. (2) Over the period 2000–2019, OCI FPs exhibited a pattern of increase, decrease, and subsequent increase. (3) OCIEI peaks in summer and reaches its lowest point in winter. Lastly, the study addresses current limitations and outlines future research objectives. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the logistic model.</p>
Full article ">Figure 2
<p>Schematic diagram of finding OCI FP using logistic modeling (Second Derivative has been enlarged to show it more clearly).</p>
Full article ">Figure 3
<p>Overview of the study area.</p>
Full article ">Figure 4
<p>R<sup>2</sup> and RMSE, the mean LST from 2000 to 2004.</p>
Full article ">Figure 5
<p>OCI FP and isotherms, the mean LST of daytime from 2000 to 2004.</p>
Full article ">Figure 6
<p>R<sup>2</sup> and RMSE, the mean LST of daytimes and seasons for 2000–2004, 2005–2009, 2010–2014, and 2015–2019.</p>
Full article ">Figure 6 Cont.
<p>R<sup>2</sup> and RMSE, the mean LST of daytimes and seasons for 2000–2004, 2005–2009, 2010–2014, and 2015–2019.</p>
Full article ">Figure 7
<p>OCI FPs for 2000–2004, 2005–2009, 2010–2014, and 2015–2019.</p>
Full article ">Figure 7 Cont.
<p>OCI FPs for 2000–2004, 2005–2009, 2010–2014, and 2015–2019.</p>
Full article ">Figure 8
<p>The Area, FRAC, and FEII of OCI FPS for different times from 2000 to 2019.</p>
Full article ">Figure 9
<p>OCIEI in different periods, 2000–2019.</p>
Full article ">Figure 10
<p>OCI FPs and land use types at different times.</p>
Full article ">
22 pages, 7918 KiB  
Article
Spatial and Temporal Change Characteristics and Climatic Drivers of Vegetation Productivity and Greenness during the 2001–2020 Growing Seasons on the Qinghai–Tibet Plateau
by Jinghan Liang, Armando Marino and Yongjie Ji
Land 2024, 13(8), 1230; https://doi.org/10.3390/land13081230 - 7 Aug 2024
Viewed by 393
Abstract
Exploring NDVI variation and what drives it on the Qinghai–Tibet Plateau can strategically inform environmental protection efforts in light of global climate change. For this analysis, we obtained MODIS NDVI data collected during the vegetative growing season, vegetation types for the region, and [...] Read more.
Exploring NDVI variation and what drives it on the Qinghai–Tibet Plateau can strategically inform environmental protection efforts in light of global climate change. For this analysis, we obtained MODIS NDVI data collected during the vegetative growing season, vegetation types for the region, and meteorological data for the same period from 2001 to 2020. We performed Theil–Sen trend analysis, Mann–Kendall significance testing, spatial autocorrelation analysis, and Hurst index calculation to review the spatiotemporal changes in NDVI characteristics on the plateau for various vegetation types. We used the correlation coefficients from these analyses to investigate how the NDVI responds to temperature and precipitation. We found the following: (1) Overall, the Qinghai–Tibet Plateau NDVI increased throughout the multi-year growing season, with a much larger area of improvement (65.68%) than of degradation (8.83%). (2) The four main vegetation types were all characterized by improvement, with meadows (72.13%) comprising the largest portion of the improved area and shrubs (18.17%) comprising the largest portion of the degraded area. (3) The spatial distribution of the NDVI had a strong positive correlation and clustering effect and was stable overall. The local clustering patterns were primarily low–low and high–high clustering. (4) The Hurst index had an average value of 0.46, indicating that the sustainability of vegetation is poor; that is, the trend of vegetation change in the growing season in a large part of the Qinghai–Tibet Plateau in the future is opposite to that in the past. (5) The plateau NDVI correlated positively with air temperature and precipitation. However, the correlations varied geographically: air temperature had a wide influence, whereas precipitation mainly influenced meadows and grassland in the northern arid zone. The overall temperature-driven effect was stronger than that of precipitation. This finding is consistent with the current research conclusion that global warming and humidification promote vegetation growth in high-altitude areas and further emphasizes the uniqueness of the Qinghai–Tibet Plateau as a climate-change-sensitive area. This study also offers a technical foundation for understanding how climate change impacts high-altitude ecosystems, as well as for formulating ecological protection strategies for the plateau. Full article
(This article belongs to the Special Issue Assessment of Land Use/Cover Change Using Geospatial Technology)
Show Figures

Figure 1

Figure 1
<p>Distribution of elevation (<b>A</b>) and vegetation types (<b>B</b>) on the Qinghai–Tibet Plateau.</p>
Full article ">Figure 2
<p>Changes in the growing-season NDVI of various vegetation types on the Qinghai–Tibet Plateau, 2001–2020.</p>
Full article ">Figure 3
<p>Spatial distribution (<b>A</b>) and significance test of trends (<b>B</b>) in NDVI average value variation during the 2001–2020 growing seasons.</p>
Full article ">Figure 4
<p>Changes in the NDVI of four major vegetation types during the 2001–2020 growing seasons on the Qinghai–Tibet Plateau.</p>
Full article ">Figure 5
<p>Significance test for NDVI change trend grades for different types of vegetation during the 2001–2020 growing seasons.</p>
Full article ">Figure 6
<p>Moran scatterplot of global NDVI spatial autocorrelation on the Qinghai–Tibet Plateau during the growing season. Note: (<b>A</b>) (2005), (<b>B</b>) (2010), (<b>C</b>) (2015), and (<b>D</b>) (2020).</p>
Full article ">Figure 7
<p>LISA map of localized NDVI spatial autocorrelation on the Qinghai–Tibet Plateau during the growing season.</p>
Full article ">Figure 8
<p>Spatial distribution of vegetation Hurst index values (<b>A</b>) and future trends (<b>B</b>) during the 2001–2020 growing seasons on the Qinghai–Tibet Plateau.</p>
Full article ">Figure 9
<p>Interannual variation diagram of precipitation (<b>A</b>) and average temperature (<b>B</b>) in the 2001–2020 growing seasons on the Qinghai–Tibet Plateau.</p>
Full article ">Figure 10
<p>Correlation of the NDVIs of the main vegetation types with precipitation and average temperature during the 2001–2020 growing seasons on the Qinghai–Tibet Plateau: (<b>A</b>) grassland, (<b>B</b>) meadow, (<b>C</b>) alpine, and (<b>D</b>) shrub).</p>
Full article ">
24 pages, 6993 KiB  
Article
Advancing Volcanic Activity Monitoring: A Near-Real-Time Approach with Remote Sensing Data Fusion for Radiative Power Estimation
by Giovanni Salvatore Di Bella, Claudia Corradino, Simona Cariello, Federica Torrisi and Ciro Del Negro
Remote Sens. 2024, 16(16), 2879; https://doi.org/10.3390/rs16162879 - 7 Aug 2024
Viewed by 527
Abstract
The global, near-real-time monitoring of volcano thermal activity has become feasible through thermal infrared sensors on various satellite platforms, which enable accurate estimations of volcanic emissions. Specifically, these sensors facilitate reliable estimation of Volcanic Radiative Power (VRP), representing the heat radiated during volcanic [...] Read more.
The global, near-real-time monitoring of volcano thermal activity has become feasible through thermal infrared sensors on various satellite platforms, which enable accurate estimations of volcanic emissions. Specifically, these sensors facilitate reliable estimation of Volcanic Radiative Power (VRP), representing the heat radiated during volcanic activity. A critical factor influencing VRP estimates is the identification of hotspots in satellite imagery, typically based on intensity. Different satellite sensors employ unique algorithms due to their distinct characteristics. Integrating data from multiple satellite sources, each with different spatial and spectral resolutions, offers a more comprehensive analysis than using individual data sources alone. We introduce an innovative Remote Sensing Data Fusion (RSDF) algorithm, developed within a Cloud Computing environment that provides scalable, on-demand computing resources and services via the internet, to monitor VRP locally using data from various multispectral satellite sensors: the polar-orbiting Moderate Resolution Imaging Spectroradiometer (MODIS), the Sea and Land Surface Temperature Radiometer (SLSTR), and the Visible Infrared Imaging Radiometer Suite (VIIRS), along with the geostationary Spinning Enhanced Visible and InfraRed Imager (SEVIRI). We describe and demonstrate the operation of this algorithm through the analysis of recent eruptive activities at the Etna and Stromboli volcanoes. The RSDF algorithm, leveraging both spatial and intensity features, demonstrates heightened sensitivity in detecting high-temperature volcanic features, thereby improving VRP monitoring compared to conventional pre-processed products available online. The overall accuracy increased significantly, with the omission rate dropping from 75.5% to 3.7% and the false detection rate decreasing from 11.0% to 4.3%. The proposed multi-sensor approach markedly enhances the ability to monitor and analyze volcanic activity. Full article
(This article belongs to the Special Issue Application of Remote Sensing Approaches in Geohazard Risk)
Show Figures

Figure 1

Figure 1
<p>(<b>1</b>) Derivation of the Normalized Thermal Index (NTI) obtained by combining the radiance of the MIR and the radiance of the TIR. (<b>2</b>) Application of the Spatial Standard Deviation (SSD) filter to each pixel in the image. (<b>3</b>) Definition of two statistical masks, Mask1 and Mask2, to identify “potential” and “true” hotspots, applied on the SSD and NTI of the volcanic area (VA). (<b>4</b>) Application of Gabor filter to extract the significant features of the image, resulting in a matrix called Gabor Weighted NTI (G-NTI). (<b>5</b>) Highlighting hotspots in the crater area and defining the Spatial Gabor Weighted NTI (SG-NTI). (<b>6</b>) Application of a statistical mask to the previously extracted matrix. (<b>7</b>) Calculation of the final VRP.</p>
Full article ">Figure 2
<p>Workflow image of the RSDF algorithm. Study cases: (<b>a</b>) Etna on 2 December 2023 at 01:10 UTC, MODIS sensor; (<b>b</b>) Etna on 15 January 2023 at 20:46 UTC, SLSTR sensor; (<b>c</b>) Stromboli on 3 October 2023 at 13:10 UTC, MODIS sensor; (<b>d</b>) Stromboli on 23 October 2023 at 09:08 UTC, SLSTR sensor.</p>
Full article ">Figure 3
<p>Time series of the Etna volcano. The panels show VRP calculated respectively from the RSDF Algorithm SLSTR (blue triangles) SLSTR Level 2 (red triangles), the RSDF Algorithm MODIS (blue triangles), MODIS Level 2 (red triangles). (<b>a</b>,<b>c</b>) shows data from January 2021 to April 2022, (<b>b</b>,<b>d</b>) from April 2022 to June 2023.</p>
Full article ">Figure 4
<p>Histograms (<b>a</b>,<b>c</b>) and probability plots (<b>b</b>,<b>d</b>) for Etna datasets. (<b>a</b>,<b>c</b>) Histograms display data distribution related to VRP (and FRP) in logarithmic scale; (<b>a</b>) blue bars represent the distribution of SLSTR-–RSDF algorithm processed data, and red bars represent SLSTR Level 2 product data; (<b>c</b>) blue bars represent the distribution of MODIS–RSDF algorithm processed data, and red bars represent MODIS active fire products. (<b>b</b>,<b>d</b>) Probability plots for normal distribution of RSDF algorithm processed data (blue), and Level 2 product data (red). The dashed grey lines represent the reference lines of the theoretical distributions, and the black dashed line in (<b>b</b>) corresponds to the slope change associated with the transition between regimes of background and high thermal activity.</p>
Full article ">Figure 5
<p>Stacked time series of VRPw (weekly mean) retrieved for the SLSTR–RSDF algorithm processed data (blue), SLSTR Level 2 product data (red), MODIS–RSDF algorithm processed data (green), and SLSTR Level 2 product data (black) at the Etna volcano, displayed on a logarithmic scale.</p>
Full article ">Figure 6
<p>VRP time series of the Stromboli volcano. The panels show VRP calculated respectively from the RSDF Algorithm SLSTR (blue triangles) SLSTR Level 2 (red triangles), the RSDF Algorithm MODIS (blue triangles), MODIS Level 2 (red triangles). (<b>a</b>,<b>c</b>) shows data from January 2021 to April 2022, (<b>b</b>,<b>d</b>) from April 2022 to June 2023.</p>
Full article ">Figure 7
<p>Histograms (<b>a</b>,<b>c</b>) and probability plots (<b>b</b>,<b>d</b>) for Stromboli datasets. (<b>a</b>,<b>c</b>) Histograms display data distribution related to VRP (and FRP) in logarithmic scale; (<b>a</b>) blue bars represent the distribution of SLSTR–RSDF algorithm processed data, and red bars represent SLSTR Level 2 product data; (<b>c</b>) blue bars represent the distribution of MODIS–RSDF algorithm processed data, and red bars represent MODIS active fire products. (<b>b</b>,<b>d</b>) Probability plots for normal distribution of RSDF algorithm processed data (blue), and Level 2 product data and MODIS active fire products (red). The dashed grey lines represent the reference lines of the theoretical distributions, and the black dashed line in (<b>b</b>) corresponds to the slope change associated with the transition between regimes of background and high thermal activity for Stromboli.</p>
Full article ">Figure 8
<p>Stacked time series of VRPw (weekly mean) retrieved for SLSTR–RSDF algorithm processed data (blue), SLSTR Level 2 product data (red), MODIS–RSDF algorithm processed data (green), and MODIS active fire products (black) at the Etna volcano, displayed on a logarithmic scale.</p>
Full article ">Figure 9
<p>Cumulative Volcanic Radiative Energy (VRE) calculated from VRP (and FRP) using the trapezoidal rule for integration. The blue line represents VRESLSTR, the red dashed line FREMODIS, the green dashed line VREMODIS, and the black dashed line FREMODIS. Panels (<b>a</b>,<b>c</b>) show data for Etna; panels (<b>b</b>,<b>d</b>) show data for Stromboli.</p>
Full article ">Figure 10
<p>Radiative power time series from SLSTR– and MODIS–RSDF algorithm data with intensity limits categorized as low, moderate, high, and extreme. (<b>a</b>) Etna, (<b>b</b>) Stromboli.</p>
Full article ">Figure 11
<p>Temporal trend of VRP values derived from the RSDF algorithm for SEVIRI, SLSTR, MODIS, and VIIRS over two periods at Mt. Etna: (<b>a</b>) 1 February 2021–30 April 2021, and (<b>b</b>) 27 September 2023–10 October 2023.</p>
Full article ">Figure 12
<p>TADR and lava flow volume flux during the effusive event at Etna from 14 May 2022 to 16 June 2022. TADR_max, TADR_mean, and TADR_min are represented by blue, red, and green points, respectively. The total volume_max, volume_mean, and volume_min are represented by blue, red, and green lines, respectively.</p>
Full article ">Figure 13
<p>TADR and lava flow volume flux during the effusive event at Stromboli from 27 September 2023 to 10 October 2023. TADR_max, TADR_mean, and TADR_min are represented by blue, red, and green points, respectively. The total volume_max, volume_mean, and volume_min are represented by blue, red, and green lines, respectively.</p>
Full article ">
21 pages, 35471 KiB  
Article
Significant Increase in African Water Vapor over 2001–2020
by Ruonan Wang, Guiping Wu, Yongwei Liu, Rong Wang, Xingwang Fan and Yuanbo Liu
Remote Sens. 2024, 16(16), 2875; https://doi.org/10.3390/rs16162875 - 6 Aug 2024
Viewed by 255
Abstract
Atmospheric water vapor is not only a key element of the global hydrological cycle but also the most abundant greenhouse gas. The phase transition and transportation of water vapor are essential for maintaining global energy balance and regulating hydrological processes. However, due to [...] Read more.
Atmospheric water vapor is not only a key element of the global hydrological cycle but also the most abundant greenhouse gas. The phase transition and transportation of water vapor are essential for maintaining global energy balance and regulating hydrological processes. However, due to insufficient meteorological observational data, climate research in Africa faces significant limitations despite its substantial contribution to changes in global precipitable water vapor (PWV). In this study, we used MODIS near-infrared (NIR) PWV products and Berkeley temperature data to depict the spatial–temporal variability in PWV across Africa from 2001 to 2020. The results reveal a significant increasing trend in PWV over Africa, with an increase of 0.0158 cm/year. Nearly 99.96% of Africa shows an increase in PWV, with 88.95% of these areas experiencing statistically significant changes, particularly in central regions of Africa. The increase in PWV is more pronounced in high-value months compared to low-value months. The equatorial region of the Congo Basin exhibits higher PWV, which gradually decreases as latitude increases. Despite significant warming (0.0162 °C/year) in Africa, there is no consistent positive correlation between temperature and water vapor. A positive relationship between PWV and temperature is observed in western Africa, while a negative relationship is noted in eastern and southern Africa on an annual scale. Additionally, an increasing trend in precipitation (4.6669 mm/year) is observed, with a significant positive correlation between PWV and precipitation across most of Africa, although this relationship varies by month. These findings provide valuable insights into the comprehension of the hydrothermal variation in Africa amidst climate warming. Full article
Show Figures

Figure 1

Figure 1
<p>The climate zones and major basins in Africa. The climate zones include tropical zone, arid zone, and temperate zone.</p>
Full article ">Figure 2
<p>Spatial distribution of (<b>a</b>) multi-year average annual and (<b>b</b>) MAM, (<b>c</b>) JJA, (<b>d</b>) SON, and (<b>e</b>) DJF seasonal PWV for 2001–2020. MAM is from March to May, JJA is from June to August, SON is from September to November, and DJF is from December to February. A, B, and C represent the climate zones of tropical, arid, and temperate, respectively.</p>
Full article ">Figure 3
<p>Spatial distribution of multi-year average of PWV in Africa from January to December for 2001–2020.</p>
Full article ">Figure 4
<p>Time series of annual average PWV in Africa for 2001–2020. The time series of the annual average PWV of (<b>a</b>) whole Africa and (<b>b</b>) climate zones (tropical, arid, and temperate climate zones). Linear trend is represented by a dashed line. (<b>c</b>) is the bar of the change rate of major basins in Africa. The red bar that represents the trend is not significant. The light blue bar and dark blue bar mean that the trend is significant at the 95% and 99% confidence levels, respectively.</p>
Full article ">Figure 5
<p>Seasonal change in PWV over Africa for 2001–2020. The time series of seasonal average of PWV over Africa (black lines), tropical zones (blue lines), arid zones (orange lines), and temperate zones (green lines) in (<b>a</b>) MAM, (<b>c</b>) JJA, (<b>e</b>) SON, and (<b>g</b>) DJF. Bar of change rate for PWV over African major basins in (<b>b</b>) MAM, (<b>d</b>) JJA, (<b>f</b>) SON, and (<b>h</b>) DJF. Dark blue bars and light blue bars represent the change rate where the levels of significance are at 0.01 and 0.05, respectively. Red bars represent that the change in PWV is not significant.</p>
Full article ">Figure 6
<p>Temporal variations in PWV in Africa on both annual and monthly scales for 2001–2020. Heatmaps for (<b>a</b>) whole of Africa, (<b>b</b>) tropical climate zones, (<b>c</b>) arid climate zones, and (<b>d</b>) temperate climate zones.</p>
Full article ">Figure 7
<p>Spatial distribution of change rate (<b>a</b>) annual and (<b>b</b>–<b>e</b>) seasonal PWV for 2001–2020. The change rate on the pixel basis is the slope of the linear regression of the time series, and the pixel of significance change (<span class="html-italic">p</span> ≤ 0.05) is marked by gray points. Dashed lines are zero and negative contours and solid lines are positive contours.</p>
Full article ">Figure 8
<p>Spatial distribution of PWV change rate from January to December for 2001–2020. The change rate on the pixel basis is the slope of the linear regression of the time series, and the pixel of significance change (<span class="html-italic">p</span> ≤ 0.05) is marked by gray points. Dashed lines are zero and negative contours and solid lines are positive contours.</p>
Full article ">Figure 9
<p>The relationship between PWV and temperature and precipitation on an annual scale for 2001–2020. (<b>a</b>) Time series of the annual average of PWV, temperature, and precipitation. Linear trend is represented by dashed line. The slope of the linear regression on the pixel basis for (<b>b</b>) PWV and temperature and (<b>d</b>) PWV and precipitation. The correlation coefficient (R<sup>2</sup>) for (<b>c</b>) PWV and temperature and (<b>e</b>) PWV and precipitation. Significant relationships (<span class="html-italic">p</span> &lt; 0.05) are marked by gray points. Dashed lines are zero and negative contours and solid lines are positive contours.</p>
Full article ">Figure 10
<p>Spatial distribution of the controlling climate indices of PWV for 2001–2020. The climate indices include TNA (Tropical Northern Atlantic Index), TSA (Tropical Southern Atlantic Index), NAO (North Atlantic Oscillation), AMO (Atlantic Multidecadal Oscillation), PDO (Pacific Decadal Oscillation), PNA (Pacific North American Index), SOI (Southern Oscillation Index), TNI (Trans-Niño Index), Niño 1.2 (Extreme Eastern Tropical Pacific Sea Surface Temperature), Niño 4 (Central Tropical Pacific Sea Surface Temperature), Niño 3.4 (East Central Tropical Pacific Sea Surface Temperature), MEI (Multivariate ENSO Index), AO (Arctic Oscillation), and AAO (Antarctic Oscillation).</p>
Full article ">Figure 11
<p>Spatial distribution of the relationship between PWV and temperature on a monthly scale for 2001–2020. (<b>a</b>–<b>f</b>) and (<b>m</b>–<b>r</b>) represent the slope of the linear regression at each pixel level, while (<b>g</b>–<b>l</b>) and (<b>s</b>–<b>x</b>) indicate the correlation coefficient between PWV and temperature at each pixel level. Grey points denote the pixel with significant relationships (<span class="html-italic">p</span> &lt; 0.05). Dashed lines represent zero and negative contours, while solid lines represent positive contours.</p>
Full article ">Figure 12
<p>Spatial distribution of the relationship between PWV and precipitation on a monthly scale for 2001–2020. (<b>a</b>–<b>f</b>) and (<b>m</b>–<b>r</b>) represent the slope of the linear regression at each pixel level, while (<b>g</b>–<b>l</b>) and (<b>s</b>–<b>x</b>) indicate the correlation coefficient between PWV and temperature at each pixel level. Grey points denote the pixel with significant relationships (<span class="html-italic">p</span> &lt; 0.05). Dashed lines represent zero and negative contours, while solid lines represent positive contours.</p>
Full article ">
25 pages, 5377 KiB  
Article
Investigating FWI Moisture Codes in Relation to Satellite-Derived Soil Moisture Data across Varied Resolutions
by Hatice Atalay, Ayse Filiz Sunar and Adalet Dervisoglu
Fire 2024, 7(8), 272; https://doi.org/10.3390/fire7080272 - 5 Aug 2024
Viewed by 389
Abstract
In the Mediterranean region, particularly in Antalya, southern Türkiye, rising forest fire risks due to climate change threaten ecosystems, property, and lives. Reduced soil moisture during the growing season is a key factor increasing fire risk by stressing plants and lowering fuel moisture [...] Read more.
In the Mediterranean region, particularly in Antalya, southern Türkiye, rising forest fire risks due to climate change threaten ecosystems, property, and lives. Reduced soil moisture during the growing season is a key factor increasing fire risk by stressing plants and lowering fuel moisture content. This study assessed soil moisture and fuel moisture content (FMC) in ten fires (2019–2021) affecting over 50 hectares. The Fire Weather Index (FWI) and its components (FFMC, DMC, DC) were calculated using data from the General Directorate of Meteorology, EFFIS (8 km), and ERA5 (≈28 km) satellite sources. Relationships between FMCs, satellite-based soil moisture datasets (SMAP, SMOS), and land surface temperature (LST) data (MODIS, Landsat 8) were analyzed. Strong correlations were found between FWI codes and satellite soil moisture, particularly with SMAP. Positive correlations were observed between LST and FWIs, while negative correlations were evident with soil moisture. Statistical models integrating in situ soil moisture and EFFIS FWI (R: −0.86, −0.84, −0.83 for FFMC, DMC, DC) predicted soil moisture levels during extended fire events effectively, with model accuracy assessed through RMSE (0.60–3.64%). The SMAP surface (0–5 cm) dataset yielded a lower RMSE of 0.60–2.08%, aligning with its higher correlation. This study underlines the critical role of soil moisture in comprehensive fire risk assessments and highlights the necessity of incorporating modeled soil moisture data in fire management strategies, particularly in regions lacking comprehensive in situ monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the study area (Antalya region); created using ArcGIS Pro (version 3.1, Esri, Redlands, CA, USA). (<b>b</b>) geographical distribution of climate types in Türkiye based on Köppen–Geiger climate system (the map is retrieved from [<a href="#B70-fire-07-00272" class="html-bibr">70</a>,<a href="#B71-fire-07-00272" class="html-bibr">71</a>]).</p>
Full article ">Figure 2
<p>Locations and sizes of ten forest fires analyzed (data sourced from EFFIS), along with meteorological stations in the study area; created using ArcGIS Pro (version 3.1, Esri, Redlands, CA, USA).</p>
Full article ">Figure 3
<p>Important meteorological parameters considered for each date in the analysis.</p>
Full article ">Figure 4
<p>Distribution of LULC classes as a percentage of burned areas in ten fires analyzed.</p>
Full article ">Figure 5
<p>Categorization of forest floor fuels based on the fuel moisture codes of the FWI System [<a href="#B28-fire-07-00272" class="html-bibr">28</a>].</p>
Full article ">Figure 6
<p>Flowchart of the study.</p>
Full article ">
12 pages, 6015 KiB  
Article
Local Evapotranspiration Is the Only Relevant Source of Moisture at the Onset of the Rainy Season in South America
by Verônica Versieux and Marcos Heil Costa
Atmosphere 2024, 15(8), 932; https://doi.org/10.3390/atmos15080932 - 4 Aug 2024
Viewed by 446
Abstract
The South American Monsoon System, which transports moisture from Amazonia to Central-West Brazil, is an important moisture source for the summer rainy season in this region. While local evapotranspiration also contributes to the atmospheric moisture supply, the balance between local and remote sources [...] Read more.
The South American Monsoon System, which transports moisture from Amazonia to Central-West Brazil, is an important moisture source for the summer rainy season in this region. While local evapotranspiration also contributes to the atmospheric moisture supply, the balance between local and remote sources during the onset of the rainy season remains uncertain. Our research aimed to quantify the role of local evapotranspiration in initiating the rainy season in Central-West Brazil. By utilizing data from various sources, such as remote sensing (MODIS), modern reanalysis (ECMWF’s ERA5), and composite products of rainfall (CHIRPS), and analyzing them in a comparative way, we conclusively found that local evapotranspiration is the only relevant source of moisture to the atmosphere during the dry-to-wet season transition, preceding the establishment of the monsoon system. Full article
(This article belongs to the Special Issue Land-Atmosphere Interactions)
Show Figures

Figure 1

Figure 1
<p>Orientation map (<b>A</b>); representation of land use and land cover in the study area (<b>B</b>) and the mean moisture flux direction in the six fortnights before and after the climatological onset of the rainy season in Mato Grosso (<b>C</b>–<b>H</b>), showing the gradual shift in patterns, from a westward direction to a southeastward direction, which characterizes the beginning of the South American monsoon.</p>
Full article ">Figure 2
<p>Comparison of evaporation data from MODIS (<b>G</b>–<b>L</b>) and ERA5 (<b>A</b>–<b>F</b>).</p>
Full article ">Figure 3
<p>Main components of the vertical water balance for Mato Grosso for six fortnights before and after the climatological onset of the rainy season in Mato Grosso. (<b>A</b>–<b>F</b>) Climatological precipitation from CHIRPS; (<b>G</b>–<b>L</b>) climatological convergence of water vapor from ERA-5 reanalysis, with negative values indicating divergence; (<b>M</b>–<b>R</b>) climatological evapotranspiration from ERA-5 reanalysis; (<b>S</b>–<b>X</b>) climatological evapotranspiration from MODIS (MYD16 product).</p>
Full article ">Figure 4
<p>Ratio between evapotranspiration and precipitation, showing the role of evapotranspiration in the formation of precipitation. (<b>A</b>–<b>F</b>) Analysis utilizing evapotranspiration from MODIS (MYD16 product); (<b>G</b>–<b>L</b>) analysis utilizing evapotranspiration from ERA-5 reanalysis. The gray area represents cells with climatological <span class="html-italic">P</span> &lt; 1 mm/day.</p>
Full article ">Figure 5
<p>Time series for all variables considering the two climate areas’ average. The latitude range of 7° S–13° S represents the forest region. The latitude range of 13° S–18° S represents the savanna subtropical region.</p>
Full article ">
22 pages, 3913 KiB  
Article
Flood Extent Delineation and Exposure Assessment in Senegal Using the Google Earth Engine: The 2022 Event
by Bocar Sy, Fatoumata Bintou Bah and Hy Dao
Water 2024, 16(15), 2201; https://doi.org/10.3390/w16152201 - 2 Aug 2024
Viewed by 684
Abstract
This study addresses the pressing need for flood extent and exposure information in data-scarce and vulnerable regions, with a specific focus on West Africa, particularly Senegal. Leveraging the Google Earth Engine (GEE) platform and integrating data from the Sentinel-1 SAR, Global Surface Water, [...] Read more.
This study addresses the pressing need for flood extent and exposure information in data-scarce and vulnerable regions, with a specific focus on West Africa, particularly Senegal. Leveraging the Google Earth Engine (GEE) platform and integrating data from the Sentinel-1 SAR, Global Surface Water, HydroSHEDS, the Global Human Settlement Layer, and MODIS land cover type, our primary objective is to delineate the extent of flooding and compare this with flooding for a one-in-a-hundred-year flood event, offering a comprehensive assessment of exposure during the period from July to October 2022 across Senegal’s 14 regions. The findings underscore a total inundation area of 2951 square kilometers, impacting 782,681 people, 238 square kilometers of urbanized area, and 21 square kilometers of farmland. Notably, August witnessed the largest flood extent, reaching 780 square kilometers, accounting for 0.40% of the country’s land area. Other regions, including Saint-Louis, Ziguinchor, Fatick, and Matam, experienced varying extents of flooding, with the data for August showing a 1.34% overlap with flooding for a one-in-a-hundred-year flood event derived from hydrological and hydraulic modeling. This low percentage reveals the distinct purpose and nature of the two approaches (remote sensing and modeling), as well as their complementarity. In terms of flood exposure, October emerges as the most critical month, affecting 281,406 people (1.56% of the population). The Dakar, Diourbel, Thiès, and Saint-Louis regions bore substantial impacts, affecting 437,025; 171,537; 115,552; and 77,501 people, respectively. These findings emphasize the imperative for comprehensive disaster preparation and mitigation efforts. This study provides a crucial national-scale perspective to guide Senegal’s authorities in formulating effective flood management, intervention, and adaptation strategies. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area. The insert in the top right corner locates our study area on the African continent. The left-hand map represents our study area, Senegal, with 14 administrative regions, and shaded relief as the map background. The 14 regions are designated by numbers 1 to 14. The corresponding names are provided in <a href="#app1-water-16-02201" class="html-app">Supplementary Table S1</a>.</p>
Full article ">Figure 2
<p>Framework for flood extent delineation and flood exposure assessment: remote sensing methodology (green), flooding for a one-in-a-hundred-year flood event methodology using modeling (blue), and methodology for estimating exposed population, urban areas, and farmland (black).</p>
Full article ">Figure 3
<p>Spatial distribution of flooded areas based on Sentinel-1, GSW, and HydroSHEDS data for the 2022 flood event per region and month: July (red), August (orange), September (yellow), and October (light green).</p>
Full article ">Figure 4
<p>Histogram of flooded areas based on Sentinel-1, GSW, and HydroSHEDS data for the 2022 flood event per region and month: July (red), August (orange), September (yellow), and October (light green).</p>
Full article ">Figure 5
<p>Spatial distribution of flooded areas based on Sentinel-1, GSW, and HydroSHEDS data for the 2022 flood event in the two most flooded regions: (<b>a</b>) Saint-Louis [<a href="#B4-water-16-02201" class="html-bibr">4</a>] (<b>top</b>) and (<b>b</b>) Ziguinchor [<a href="#B2-water-16-02201" class="html-bibr">2</a>] (<b>bottom</b>).</p>
Full article ">Figure 6
<p>Population exposed to flooding from July to October 2022, estimated using the intersection of GHLS population datasets with the flooded areas in the Google Earth Engine.</p>
Full article ">Figure 7
<p>Spatial distribution of the two most exposed regions to flooding by population: (<b>a</b>) Dakar and (<b>b</b>) Diourbel. Assessed through the intersection of GHLS population datasets with the flooded areas in the Google Earth Engine.</p>
Full article ">Figure 7 Cont.
<p>Spatial distribution of the two most exposed regions to flooding by population: (<b>a</b>) Dakar and (<b>b</b>) Diourbel. Assessed through the intersection of GHLS population datasets with the flooded areas in the Google Earth Engine.</p>
Full article ">Figure 8
<p>(<b>a</b>) Urban areas and (<b>b</b>) farmland exposed to flooding, derived from the intersection of MODIS land cover datasets with the flooded areas in the Google Earth Engine.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>) Urban areas and (<b>b</b>) farmland exposed to flooding, derived from the intersection of MODIS land cover datasets with the flooded areas in the Google Earth Engine.</p>
Full article ">
18 pages, 7349 KiB  
Article
Temporal Patterns of Vegetation Greenness for the Main Forest-Forming Tree Species in the European Temperate Zone
by Kinga Kulesza and Agata Hościło
Remote Sens. 2024, 16(15), 2844; https://doi.org/10.3390/rs16152844 - 2 Aug 2024
Viewed by 301
Abstract
In light of recently accelerating global warming, the changes in vegetation trends are vital for the monitoring of the dynamics of both whole ecosystems and individual species. Detecting changes within the time series of specific forest ecosystems or species is very important in [...] Read more.
In light of recently accelerating global warming, the changes in vegetation trends are vital for the monitoring of the dynamics of both whole ecosystems and individual species. Detecting changes within the time series of specific forest ecosystems or species is very important in the context of assessing their vulnerability to climate change and other negative phenomena. Hence, the aim of this paper was to identify the trend change points and periods of greening and browning in multi-annual time series of the normalised difference vegetation index (NDVI) and enhanced vegetation index (EVI) of four main forest-forming tree species in the temperate zone: pine, spruce, oak and beech. The research was conducted over the last two decades (2002–2022), and was based on vegetation indices data derived from the Moderate Resolution Imaging Spectroradiometer (MODIS). To this end, several research approaches, including calculating the linear trends in the moving periods and BEAST algorithm, were adapted. A pattern of browning then greening then constant was detected for coniferous species, mostly pine. In turn, for broadleaved species, namely oak and beech, a pattern of greening then constant was identified, without the initial phase of browning. The main trend change points seem to be ca. 2006 and ca. 2015 for coniferous species and solely around 2015 for deciduous ones. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the regional directorates of state forests used in this study (1—Wroclaw, 2—Lodz, 3—Lublin) and spatial distribution of the tree species masks in MODIS grid (orange—pine, blue—spruce, purple—oak, green—beech). In the upper right-hand corner the directorates’ borders are presented over the elevation map (source of the elevation map: <a href="http://pl.wikipedia.org" target="_blank">pl.wikipedia.org</a>, accessed on 21 October 2022). The black dashed line indicates the pine-forested area showcased in detail in this study.</p>
Full article ">Figure 2
<p>The slopes (z-score for NDVI·month<sup>−1</sup>) of the linear trends of monthly NDVI z-scores over four species (pine, spruce, oak and beech) (black lines) and their <span class="html-italic">p</span>-values (blue lines) from 2002 to 2022 in moving 5-year (60-month) periods. The statistical significance level of α = 0.05 is indicated by blue dashed lines. Statistically significant slopes are additionally highlighted with a grey colour.</p>
Full article ">Figure 3
<p>The slopes (z-score for EVI·month<sup>−1</sup>) of the linear trends of monthly EVI z-scores over four species (pine, spruce, oak and beech) (black lines) and their <span class="html-italic">p</span>-values (blue lines) from 2002 to 2022 in moving 5-year (60-month) periods. The statistical significance level of α = 0.05 is indicated by blue dashed lines. Statistically significant slopes are additionally highlighted with a grey colour.</p>
Full article ">Figure 4
<p>The course of NDVI z-scores (blue line) for four tree species (pine, spruce, oak and beech) in the period 2002–2022. The graphs also present the running 37-month mean (orange line) and linear trends for the shorter periods (coloured are lines explained next to each graph), together with trend line equations, coefficients of determination <span class="html-italic">R</span><sup>2</sup> and <span class="html-italic">p</span>-values.</p>
Full article ">Figure 5
<p>The course of EVI z-scores (blue line) for four tree species (pine, spruce, oak and beech) in the period 2002–2022. The graphs also present the running 37-month mean (orange line) and linear trends for the shorter periods (coloured lines are explained next to each graph), together with trend line equations, coefficients of determination <span class="html-italic">R</span><sup>2</sup> and <span class="html-italic">p</span>-values.</p>
Full article ">Figure 6
<p>Slopes of the trends in NDVI z-scores (z-score for NDVI·month<sup>−1</sup>) in a selected pine-forested area in regional directorate of state forests in Wroclaw (location of this area is presented in <a href="#remotesensing-16-02844-f001" class="html-fig">Figure 1</a>), in the periods 01.2002–12.2005 (<b>a</b>), 01.2006–12.2015 (<b>b</b>), 01.2016–06.2019 (<b>c</b>) and 07.2019–12.2022 (<b>d</b>). Grey colour indicates insignificant slopes at the significance level of α = 0.05.</p>
Full article ">Figure 7
<p>The slopes (z-score·month<sup>−1</sup>) of the linear trends of monthly z-scores for T (<b>a</b>) and P (<b>b</b>) and their <span class="html-italic">p</span>-values (blue lines) from 2002 to 2022 in moving 5-year (60-month) periods. The statistical significance level of α = 0.05 is indicated by blue dashed lines. Statistically significant slopes are additionally highlighted with a grey colour.</p>
Full article ">Figure 8
<p>The course of z-scores (blue lines) for T (<b>a</b>) and P (<b>b</b>) in the period 2002–2022. The graphs also present the running 37-month mean (orange line) and linear trends for the shorter periods (coloured lines are explained next to each graph), together with trend line equations, coefficients of determination <span class="html-italic">R</span><sup>2</sup> and <span class="html-italic">p</span>-values.</p>
Full article ">Figure 9
<p>Relationship between z-scores for T (<b>a</b>,<b>b</b>) and P (<b>c</b>,<b>d</b>) and z-scores for NDVI (<b>a</b>,<b>c</b>) and EVI (<b>b</b>,<b>d</b>) over four species (pine, spruce, oak and beech) from 2002 to 2022. Each graph consist of 1008 points (252 points for each species), and also presents linear regression (black line), together with regression equation, coefficient of determination <span class="html-italic">R</span><sup>2</sup> and <span class="html-italic">p</span>-value.</p>
Full article ">
28 pages, 20313 KiB  
Article
SHAP-Driven Explainable Artificial Intelligence Framework for Wildfire Susceptibility Mapping Using MODIS Active Fire Pixels: An In-Depth Interpretation of Contributing Factors in Izmir, Türkiye
by Muzaffer Can Iban and Oktay Aksu
Remote Sens. 2024, 16(15), 2842; https://doi.org/10.3390/rs16152842 - 2 Aug 2024
Viewed by 590
Abstract
Wildfire susceptibility maps play a crucial role in preemptively identifying regions at risk of future fires and informing decisions related to wildfire management, thereby aiding in mitigating the risks and potential damage posed by wildfires. This study employs eXplainable Artificial Intelligence (XAI) techniques, [...] Read more.
Wildfire susceptibility maps play a crucial role in preemptively identifying regions at risk of future fires and informing decisions related to wildfire management, thereby aiding in mitigating the risks and potential damage posed by wildfires. This study employs eXplainable Artificial Intelligence (XAI) techniques, particularly SHapley Additive exPlanations (SHAP), to map wildfire susceptibility in Izmir Province, Türkiye. Incorporating fifteen conditioning factors spanning topography, climate, anthropogenic influences, and vegetation characteristics, machine learning (ML) models (Random Forest, XGBoost, LightGBM) were used to predict wildfire-prone areas using freely available active fire pixel data (MODIS Active Fire Collection 6 MCD14ML product). The evaluation of the trained ML models showed that the Random Forest (RF) model outperformed XGBoost and LightGBM, achieving the highest test accuracy (95.6%). All of the classifiers demonstrated a strong predictive performance, but RF excelled in sensitivity, specificity, precision, and F-1 score, making it the preferred model for generating a wildfire susceptibility map and conducting a SHAP analysis. Unlike prevailing approaches focusing solely on global feature importance, this study fills a critical gap by employing a SHAP summary and dependence plots to comprehensively assess each factor’s contribution, enhancing the explainability and reliability of the results. The analysis reveals clear associations between factors such as wind speed, temperature, NDVI, slope, and distance to villages with increased fire susceptibility, while rainfall and distance to streams exhibit nuanced effects. The spatial distribution of the wildfire susceptibility classes highlights critical areas, particularly in flat and coastal regions near settlements and agricultural lands, emphasizing the need for enhanced awareness and preventive measures. These insights inform targeted fire management strategies, highlighting the importance of tailored interventions like firebreaks and vegetation management. However, challenges remain, including ensuring the selected factors’ adequacy across diverse regions, addressing potential biases from resampling spatially varied data, and refining the model for broader applicability. Full article
(This article belongs to the Special Issue Artificial Intelligence for Natural Hazards (AI4NH))
Show Figures

Figure 1

Figure 1
<p>Study area and wildfire inventory.</p>
Full article ">Figure 2
<p>Topographical conditioning factors: (<b>A</b>) elevation, (<b>B</b>) slope, (<b>C</b>) aspect, (<b>D</b>) TWI.</p>
Full article ">Figure 3
<p>Climatic conditioning factors: (<b>A</b>) annual average temperature, (<b>B</b>) annual rainfall, (<b>C</b>) annual solar radiation, (<b>D</b>) annual average wind speed.</p>
Full article ">Figure 4
<p>Anthropogenic conditioning factors: (<b>A</b>) LULC, (<b>B</b>) distance to roads, (<b>C</b>) distance to villages, (<b>D</b>) distance to streams.</p>
Full article ">Figure 5
<p>Vegetation-related conditioning factors: (<b>A</b>) forest type, (<b>B</b>) tree cover density, (<b>C</b>) NDVI.</p>
Full article ">Figure 6
<p>Summary of the research methodology steps utilized in the study.</p>
Full article ">Figure 7
<p>Multicollinearity test results (<b>upper</b>) before recursive elimination and (<b>lower</b>) after recursive elimination and selected factors.</p>
Full article ">Figure 8
<p>Pearson’s correlation coefficient matrix.</p>
Full article ">Figure 9
<p>Confusion matrixes.</p>
Full article ">Figure 10
<p>Classifiers’ performance comparison.</p>
Full article ">Figure 11
<p>ROC curves.</p>
Full article ">Figure 12
<p>SHAP summary plot of RF classifier’s output.</p>
Full article ">Figure 13
<p>Global feature importance by absolute SHAP values.</p>
Full article ">Figure 14
<p>Each factor’s SHAP dependence plots.</p>
Full article ">Figure 15
<p>Generated wildfire susceptibility map for the province of Izmir.</p>
Full article ">Figure 16
<p>Area extent of susceptibility classes and number of wildfire samples corresponding to each susceptibility class.</p>
Full article ">
Back to TopTop