[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (439)

Search Parameters:
Keywords = TERRA MODIS

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 32897 KiB  
Article
On the Suitability of Different Satellite Land Surface Temperature Products to Study Surface Urban Heat Islands
by Alexandra Hurduc, Sofia L. Ermida and Carlos C. DaCamara
Remote Sens. 2024, 16(20), 3765; https://doi.org/10.3390/rs16203765 - 10 Oct 2024
Viewed by 678
Abstract
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are [...] Read more.
Remote sensing satellite data have been a crucial tool in understanding urban climates. The variety of sensors with different spatiotemporal characteristics and retrieval methodologies gave rise to a multitude of approaches when analyzing the surface urban heat island effect (SUHI). Although there are considerable advantages that arise from these different characteristics (spatiotemporal resolution, time of observation, etc.), it also means that there is a need for understanding the ability of sensors in capturing spatial and temporal SUHI patterns. For this, several land surface temperature products are compared for the cities of Madrid and Paris, retrieved from five sensors: the Spinning Enhanced Visible and InfraRed Imager onboard Meteosat Second Generation, the Advanced Very-High-Resolution Radiometer onboard Metop, the Moderate-resolution Imaging Spectroradiometer onboard both Aqua and Terra, and the Thermal Infrared Sensor onboard Landsat 8 and 9. These products span a wide range of LST algorithms, including split-window, single-channel, and temperature–emissivity separation methods. Results show that the diurnal amplitude of SUHI may not be well represented when considering daytime and nighttime polar orbiting platforms. Also, significant differences arise in SUHI intensity and spatial and temporal variability due to the different methods implemented for LST retrieval. Full article
(This article belongs to the Section AI Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Land cover resampled for the three projections of LST products (<b>a</b>–<b>f</b>) along with the percentage of urban pixels (<b>g</b>–<b>l</b>).</p>
Full article ">Figure 2
<p>Time of observation of each sensor: (<b>a</b>) for Madrid during daytime and time of minimum SUHI (SUHI<sub>min</sub>), (<b>b</b>) for Madrid during nighttime and time of maximum SUHI (SUHI<sub>max</sub>), (<b>c</b>) for Paris during daytime and SUHI<sub>max</sub>, (<b>d</b>) for Paris during nighttime and SUHI<sub>min</sub>. Colored bins are sampled every 15 min.</p>
Full article ">Figure 3
<p>Mean DJF (December, January, February) LST for all products considered. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 4
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for JJA (June, July, and August). (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST maximum and SUHI minimum are shown; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST minimum and SUHI maximum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 5
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris. (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime. Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 6
<p>As <a href="#remotesensing-16-03765-f003" class="html-fig">Figure 3</a> but for Paris and DJF; (<b>a1</b>–<b>a9</b>) The spatial pattern during daytime, and in the case of the MLST, the most frequent hour of the LST and SUHI maximum; (<b>b1</b>–<b>b9</b>) histograms of urban and rural LST shown in (<b>a1</b>–<b>a9</b>); (<b>c1</b>–<b>c7</b>) as in the first line but for nighttime and for the LST and SUHI minimum; (<b>d1</b>–<b>d7</b>) as in (<b>b1</b>–<b>b9</b>) but for nighttime, an extension of the histogram in (<b>d6</b>) is seen in (<b>d8</b>). Please note that color bars are different amongst the different products to allow a better visualization of patterns, but value ranges of the histograms are the same.</p>
Full article ">Figure 7
<p>Diurnal cycle of SUHI for Madrid: (<b>a</b>) DJF, (<b>b</b>) MAM, (<b>c</b>) JJA, (<b>d</b>) SON.</p>
Full article ">Figure 8
<p>As <a href="#remotesensing-16-03765-f007" class="html-fig">Figure 7</a> but for Paris.</p>
Full article ">Figure 9
<p>Correlation of monthly SUHI anomalies between all products considered: (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</p>
Full article ">Figure 10
<p>As <a href="#remotesensing-16-03765-f009" class="html-fig">Figure 9</a> but for Paris. (<b>a</b>) daytime, (<b>b</b>) nighttime. Blank spaces correspond to pairs of products with no significant correlation (<span class="html-italic">p</span>-value &gt; 0.05).</p>
Full article ">
23 pages, 8867 KiB  
Article
Synergistic Potential of Optical and Radar Remote Sensing for Snow Cover Monitoring
by Jose-David Hidalgo-Hidalgo, Antonio-Juan Collados-Lara, David Pulido-Velazquez, Steven R. Fassnacht and C. Husillos
Remote Sens. 2024, 16(19), 3705; https://doi.org/10.3390/rs16193705 - 5 Oct 2024
Viewed by 696
Abstract
This research studies the characteristics of snow-covered area (SCA) from two vastly different sensors: optical (Moderate-Resolution Imaging Spectroradiometer, or MODIS, equipped on board the Terra satellite) and radar (Synthetic Aperture Radar (SAR) on-board Sentinel-1 satellites). The focus are the five mountain ranges of [...] Read more.
This research studies the characteristics of snow-covered area (SCA) from two vastly different sensors: optical (Moderate-Resolution Imaging Spectroradiometer, or MODIS, equipped on board the Terra satellite) and radar (Synthetic Aperture Radar (SAR) on-board Sentinel-1 satellites). The focus are the five mountain ranges of the Iberian Peninsula (Cantabrian System, Central System, Iberian Range, Pyrenees, and Sierra Nevada). The MODIS product was selected to identify SCA dynamics in these ranges using the Probability of Snow Cover Presence Index (PSCPI). In addition, we evaluate the potential advantage of the use of SAR remote sensing to complete optical SCA under cloudy conditions. For this purpose, we utilize the Copernicus High-Resolution Snow and Ice SAR Wet Snow (HRS&I SWS) product. The Pyrenees and the Sierra Nevada showed longer-lasting SCA duration and a higher PSCPI throughout the average year. Moreover, we demonstrate that the latitude gradient has a significant influence on the snowline elevation in the Iberian mountains (R2 ≥ 0.84). In the Iberian mountains, a general negative SCA trend is observed due to the recent climate change impacts, with a particularly pronounced decline in the winter months (December and January). Finally, in the Pyrenees, we found that wet snow detection has high potential for the spatial gap-filling of MODIS SCA in spring, contributing above 27% to the total SCA. Notably, the additional SCA provided in winter is also significant. Based on the results obtained in the Pyrenees, we can conclude that implementing techniques that combine SAR and optical satellite sensors for SCA detection may provide valuable additional SCA data for the other Iberian mountains, in which the radar product is not available. Full article
(This article belongs to the Special Issue Multi-Source Remote Sensing Data in Hydrology and Water Management)
Show Figures

Figure 1

Figure 1
<p>Location of the main snow-dominated mountain ranges of the Iberian Peninsula. The circle points indicate the centroid of each mountain range.</p>
Full article ">Figure 2
<p>Flowchart of the methodology.</p>
Full article ">Figure 3
<p>Aggregated climate variables sourced from the AEMET 5 km dataset to appraise the differences between Iberian montane regions over an average year for (<b>a</b>) mean monthly accumulated precipitation and (<b>b</b>) minimum, average, and maximum temperature.</p>
Full article ">Figure 4
<p>Distribution of the Probability of Snow Cover Presence Index (PSCPI) per elevation over the main mountain ranges of the Iberian Peninsula: (<b>a</b>) Sierra Nevada, (<b>b</b>) Pyrenees, (<b>c</b>) Cantabrian System, (<b>d</b>) Iberian Range and (<b>e</b>) Central System. The red color scatterplots indicate the average PSCPI per elevation. The shaded area represents PSCPI values within the confidence interval between 5% and 95%.</p>
Full article ">Figure 5
<p>Daily distribution of the Probability of Snow Cover Presence Index (PSCPI) over the main Iberian mountains taking into account different elevation ranges. The right numbers indicate the percentage of area relative to the total area corresponding to each elevation band.</p>
Full article ">Figure 6
<p>(<b>a</b>) Linear correlation between latitude and snowline elevation based on different thresholds (THR) of the average PSCPI over the main snow-dominated Iberian mountains; (<b>b</b>) Determination coefficient of the simple linear regression between latitude and snowline elevation associated with the examined average PSCPI thresholds.</p>
Full article ">Figure 7
<p>Five-year period temporal evolution of SCA in terms of the average PSCPI for the entire domain of the mountain ranges of the Iberian Peninsula on a monthly basis.</p>
Full article ">Figure 8
<p>(<b>a</b>) Monthly PWSAI distribution and contribution of MODIS and HRS&amp;I SWS (wet snow) products in relation to the total SCA over an average year in the Pyrenees. (<b>b</b>) Monthly variation of the contribution of HRS&amp;I SWS product to the total SCA over an average year in the Pyrenees. The red line indicates the median, and the edges of the box (blue color) represent the first quartile (bottom edge) and third quartile (top edge). The upper adjacent is the furthest observation within one and a half times the interquartile range of the lower end of the box, and the upper adjacent is the furthest observation within one and a half times the interquartile range of the upper end of the box. Outliers are considered as the values greater than the upper adjacent.</p>
Full article ">Figure 9
<p>Examples of the combination of SAR (HRS&amp;I SWS) and optical (MODIS) remote sensing products for the following dates: 04-February-2017 (winter) and 18-April-2018 (spring).</p>
Full article ">Figure A1
<p>Monthly distribution of the Probability of Cloud Presence Index (PCPI) for an average year over the main Iberian mountains.</p>
Full article ">Figure A2
<p>Distribution of the tree-cover density (TCD) per elevation over the main mountain ranges of the Iberian Peninsula: (<b>a</b>) Sierra Nevada, (<b>b</b>) Pyrenees, (<b>c</b>) Cantabrian System, (<b>d</b>) Iberian Range, and (<b>e</b>) Central System. The red color scatterplots indicate the average TCD per elevation. The shaded area represents TCD values within the confidence interval between 10% and 90%.</p>
Full article ">Figure A2 Cont.
<p>Distribution of the tree-cover density (TCD) per elevation over the main mountain ranges of the Iberian Peninsula: (<b>a</b>) Sierra Nevada, (<b>b</b>) Pyrenees, (<b>c</b>) Cantabrian System, (<b>d</b>) Iberian Range, and (<b>e</b>) Central System. The red color scatterplots indicate the average TCD per elevation. The shaded area represents TCD values within the confidence interval between 10% and 90%.</p>
Full article ">
20 pages, 16133 KiB  
Article
Changes in Vegetation Cover and the Relationship with Surface Temperature in the Cananéia–Iguape Coastal System, São Paulo, Brazil
by Jakeline Baratto, Paulo Miguel de Bodas Terassi and Emerson Galvani
Remote Sens. 2024, 16(18), 3460; https://doi.org/10.3390/rs16183460 - 18 Sep 2024
Viewed by 622
Abstract
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized [...] Read more.
The objective of this article is to investigate the possible correlations between vegetation indices and surface temperature in the Cananéia–Iguape Coastal System (CICS), in São Paulo (Brazil). Vegetation index data from MODIS orbital products were used to carry out this work. The Normalized Difference Vegetation Index (NDVI) and the Enhanced Vegetation Index (EVI) were acquired from the MODIS/Aqua sensor (MYD13Q1) and the leaf area index (LAI) from the MODIS/Terra (MOD15A2H). Surface temperature data were acquired from MODIS/Aqua (MYD11A2). The data were processed using Google Earth Engine and Google Colab. The data were collected, and spatial and temporal correlations were applied. Correlations were applied in the annual and seasonal period. The annual temporal correlation between vegetation indices and surface temperature was positive, but statistically significant for the LAI, with r = 0.43 (90% significance). In the seasonal period, positive correlations occurred in JFM for all indices (95% significance). Spatially, the results of this research indicate that the largest area showed a positive correlation between VI and LST. The hottest and rainiest periods (OND and JFM) had clearer and more significant correlations. In some regions, significant and clear correlations were observed, such as in some areas in the north, south and close to the city of Iguape. This highlights the complexity of the interactions between vegetation indices and climatic attributes, and highlights the importance of considering other environmental variables and processes when interpreting changes in vegetation. However, this research has significantly progressed the field, by establishing new correlations and demonstrating the importance of considering climate variability, for a more accurate understanding of the impacts on vegetation indices. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area (<b>A</b>,<b>B</b>) and land use mapping (<b>C</b>).</p>
Full article ">Figure 2
<p>Variation in surface temperature and monthly (<b>A</b>) and annual (<b>B</b>) rainfall for the Cananéia-Iguape Coastal System for the 20032022 period. Source: MODIS/Aqua and CHIRPS, 2024.</p>
Full article ">Figure 3
<p>Annual variation in vegetation indices for the 2003–2022 period in the Cananéia–Iguape Coastal System.</p>
Full article ">Figure 4
<p>Scatter plot of annual NDVI (<b>a</b>), EVI (<b>b</b>) and LAI (<b>c</b>) values and surface temperature from 2003 to 2022.</p>
Full article ">Figure 5
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and surface temperature for the JFM (<b>a</b>–<b>c</b>) and AMJ (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 6
<p>Scatter plot of seasonal values of VI–NDVI (<b>a</b>,<b>d</b>), EVI (<b>b</b>,<b>e</b>) and LAI (<b>c</b>,<b>f</b>)—and climate variables for the JAS (<b>a</b>–<b>c</b>) and OND (<b>d</b>–<b>f</b>) quarter.</p>
Full article ">Figure 7
<p>Annual linear correlation between surface temperature and NDVI (<b>A</b>), EVI (<b>B</b>) and LAI (<b>C</b>) between 2003 and 2022.</p>
Full article ">Figure 8
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JFM (<b>A</b>–<b>C</b>) and AMJ (<b>D</b>–<b>F</b>) periods.</p>
Full article ">Figure 9
<p>Seasonal linear correlation between surface temperature and VI between 2004 and 2022 for the JAS (<b>A</b>–<b>C</b>) and OND (<b>D</b>–<b>F</b>) periods.</p>
Full article ">
17 pages, 34922 KiB  
Article
Coastal Sea Ice Concentration Derived from Marine Radar Images: A Case Study from Utqiaġvik, Alaska
by Felix St-Denis, L. Bruno Tremblay, Andrew R. Mahoney and Kitrea Pacifica L. M. Takata-Glushkoff
Remote Sens. 2024, 16(18), 3357; https://doi.org/10.3390/rs16183357 - 10 Sep 2024
Viewed by 597
Abstract
We apply the Canny edge algorithm to imagery from the Utqiaġvik coastal sea ice radar system (CSIRS) to identify regions of open water and sea ice and quantify ice concentration. The radar-derived sea ice concentration (SIC) is compared against the (closest to the [...] Read more.
We apply the Canny edge algorithm to imagery from the Utqiaġvik coastal sea ice radar system (CSIRS) to identify regions of open water and sea ice and quantify ice concentration. The radar-derived sea ice concentration (SIC) is compared against the (closest to the radar field of view) 25 km resolution NSIDC Climate Data Record (CDR) and the 1 km merged MODIS-AMSR2 sea ice concentrations within the ∼11 km field of view for the year 2022–2023, when improved image contrast was first implemented. The algorithm was first optimized using sea ice concentration from 14 different images and 10 ice analysts (140 analyses in total) covering a range of ice conditions with landfast ice, drifting ice, and open water. The algorithm is also validated quantitatively against high-resolution MODIS-Terra in the visible range. Results show a correlation coefficient and mean bias error between the optimized algorithm, the CDR and MODIS-AMSR2 daily SIC of 0.18 and 0.54, and ∼−1.0 and 0.7%, respectively, with an averaged inter-analyst error of ±3%. In general, the CDR captures the melt period correctly and overestimates the SIC during the winter and freeze-up period, while the merged MODIS-AMSR2 better captures the punctual break-out events in winter, including those during the freeze-up events (reduction in SIC). Remnant issues with the detection algorithm include the false detection of sea ice in the presence of fog or precipitation (up to 20%), quantified from the summer reconstruction with known open water conditions. The proposed technique allows for the derivation of the SIC from CSIRS data at spatial and temporal scales that coincide with those at which coastal communities members interact with sea ice. Moreover, by measuring the SIC in nearshore waters adjacent to the shoreline, we can quantify the effect of land contamination that detracts from the usefulness of satellite-derived SIC for coastal communities. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the study site and location of the CSIRS. The black circle marks the coastal radar range and the black start highlights the location of the coastal radar.</p>
Full article ">Figure 2
<p>Flow chart of the floe edge detection algorithm.</p>
Full article ">Figure 3
<p>Images from 11 March 2022, taken after each algorithm step: (<b>a</b>) initial image, (<b>b</b>) image with land removed, (<b>c</b>) the output of the Canny edge algorithm, (<b>d</b>) the contours, in red, found from the detected edges, and (<b>e</b>) the final sea ice contour.</p>
Full article ">Figure 4
<p>Images analyzed by the analysts.</p>
Full article ">Figure 5
<p>CDR (<b>a</b>) and merged MODIS-AMSR2 (<b>b</b>) grid cells used for the comparison with the marine radar. Utqiaġvik and the radar range are marked with a black star and circle.</p>
Full article ">Figure 6
<p>(<b>a</b>) Minimum RMSE (red) and the corresponding best 1:1 line fit (blue) for each kernel, (<b>b</b>) scatter plot of SIC derived from the radar images with the optimal set of parameters and SIC from the 10 analysts (colorbar), including the best-line fit (green line), (<b>c</b>) the analyst standard deviation (STD) for each analyzed frame, and (<b>d</b>) histogram of the departure of the best 1:1 line fit when removing one-by-one the analyzed frame. Note that the blue axis in (<b>a</b>) does not start at 0. Each of the horizontal lines represent a different image in (<b>b</b>). The inter-analyst averaged error of 0.026 is represented by the dashed line in (<b>c</b>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Nearly synchronous marine radar image, including sea ice edge (red) from the detection algorithm and at 15:09 local time (AKDT). (<b>b</b>) MODIS Terra image at 15:09 AKDT. (<b>c</b>) Time series of Pearson correlation coefficient (r) between all coarse-grained (1 km) 4 min images (360 in total) CSIRS SIC and the merged MODIS-AMSR2 SIC for 14 April 2022. The red shading corresponds to the 95% confidence interval. (<b>d</b>) Scatter plot of the CSIRS SIC with the merged MODIS-AMSR2 for the time of maximum correlation at 6:00 AKDT. The best line fit is given by <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0.43</mn> <mi>x</mi> <mo>+</mo> <mn>0.28</mn> </mrow> </semantics></math> and the red shading corresponds to the RMSE of 0.12.</p>
Full article ">Figure 8
<p>Daily (<b>a</b>), 7-day running mean (<b>b</b>), and 31-day running mean (<b>c</b>) time series of the SIC derived from the radar (blue), the CDR (green), and the merged MODIS-AMSR2 (red) for 2022 to 2023 as a function of the Julian days starting 1 January 2022. The holes in the time series represent the non-availability of the data.</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 783
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 ">
20 pages, 6363 KiB  
Article
Temporal-Difference Graph-Based Optimization for High-Quality Reconstruction of MODIS NDVI Data
by Shengtai Ji, Shuxin Han, Jiaxin Hu, Yuguang Li and Jing-Cheng Han
Remote Sens. 2024, 16(15), 2713; https://doi.org/10.3390/rs16152713 - 24 Jul 2024
Viewed by 777
Abstract
The Normalized Difference Vegetation Index (NDVI) is a crucial remote-sensing metric for assessing land surface vegetation greenness, essential for various studies encompassing phenology, ecology, hydrology, etc. However, effective applications of NDVI data are hindered by data noise due to factors such as cloud [...] Read more.
The Normalized Difference Vegetation Index (NDVI) is a crucial remote-sensing metric for assessing land surface vegetation greenness, essential for various studies encompassing phenology, ecology, hydrology, etc. However, effective applications of NDVI data are hindered by data noise due to factors such as cloud contamination, posing challenges for accurate observation. In this study, we proposed a novel approach for employing a Temporal-Difference Graph (TDG) method to reconstruct low-quality pixels in NDVI data. Regarding spatio-temporal NDVI data as a time-varying graph signal, the developed method utilized an optimization algorithm to maximize the spatial smoothness of temporal differences while preserving the spatial NDVI pattern. This approach was further evaluated by reconstructing MODIS/Terra Vegetation Indices 16-Day L3 Global 250 m Grid (MOD13Q1) products over Northwest China. Through quantitative comparison with a previous state-of-the-art method, the Savitzky–Golay (SG) filter method, the obtained results demonstrated the superior performance of the TDG method, and highly accurate results were achieved in both the temporal and spatial domains irrespective of noise types (positively-biased, negatively-biased, or linearly-interpolated noise). In addition, the TDG-based optimization approach shows great robustness to noise intensity within spatio-temporal NDVI data, suggesting promising prospects for its application to similar datasets. Full article
(This article belongs to the Section Environmental Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Flowchart of research methodology.</p>
Full article ">Figure 2
<p>Schematic illustration of a <span class="html-italic">k</span>-nearest-neighbor rule to establish edges (<span class="html-italic">k</span> = 4 in (<b>a</b>) and <span class="html-italic">k</span> = 8 in (<b>b</b>)). For each vertex (black circle), only the top <span class="html-italic">k</span> nearest neighboring vertices (gray filled circles) are connected by an edge.</p>
Full article ">Figure 3
<p>Location of the study area, showing land cover types over this region. The red points refer to the selected sample patches on which NDVI reconstruction methods are tested.</p>
Full article ">Figure 4
<p>Comparison of the time series of original NDVI (red dotted lines), reconstructed NDVI using the TDG method (blue lines), and reconstructed NDVI using the SG filter method (green lines). The original NDVI data are contaminated using (<b>a</b>) PM, (<b>b</b>) NM, and (<b>c</b>) ND artificial noise. Background colors indicate the PRI categories of the corresponding entries where white indicates “good data”; light gray indicates “marginal data”; dark gray indicates “snow/ice”, “cloudy”, or “fill/no data”; and orange indicates noise-introduced entries that were “good data” originally but were modified to (<b>a</b>,<b>b</b>) “marginal data” or (<b>c</b>) “fill/no data”.</p>
Full article ">Figure 5
<p>Four sets of entire original NDVI time series (red dotted lines), reconstructed NDVI using the TDG method (blue lines), and reconstructed NDVI using the SG filter method (green lines) without introducing artificial noise. The good-data rate along the time series is (<b>a</b>) 76.92%, (<b>b</b>) 45.13%, (<b>c</b>) 24.10%, or (<b>d</b>) 19.23%.</p>
Full article ">Figure 6
<p>Four sets of NDVI spatial profiles showing the PRI of the corresponding profiles and comparing between the original NDVI, the reconstructed NDVI using the SG filter method, and the reconstructed NDVI using the TDG filter method. No artificial noise is introduced. The spatial profiles are (<b>a</b>) Patch 6 on 28 July 2009, (<b>b</b>) Patch 3 on 21 March 2004, (<b>c</b>) Patch 7 on 26 June 2010, and (<b>d</b>) Patch 7 on 15 October 2004. The good-data rate of the profile is (<b>a</b>) 96.22%, (<b>b</b>) 45.43%, (<b>c</b>) 0%, or (<b>d</b>) 0%. Some zoomed-in profiles over small regions (30 × 30 pixels) are illustrated in (<b>e</b>) to provide clearer comparison of the reconstruction results using the two methods.</p>
Full article ">Figure 7
<p>Variation in the RMSE of reconstructed NDVI with respect to the good-data rate of the given NDVI data in Patch 6. NDVI data are reconstructed using the TDG method or the SG filter method.</p>
Full article ">
15 pages, 6472 KiB  
Article
Spatio-Temporal Behavior of Land Surface Temperatures (LSTs) in Central Chile, Using Terra MODIS Images
by Pedro Muñoz-Aguayo, Luis Morales-Salinas, Roberto Pizarro, Alfredo Ibáñez, Claudia Sangüesa, Guillermo Fuentes-Jaque, Cristóbal Toledo and Pablo A. Garcia-Chevesich
Hydrology 2024, 11(7), 103; https://doi.org/10.3390/hydrology11070103 - 12 Jul 2024
Viewed by 1434
Abstract
Land surface temperature (LST) is one of the most important variables in the physical processes of surface energy and water balance. The temporal behavior of LST was analyzed between the latitudes 32°00′ S and 34°24′ S (Valparaíso and Metropolitana regions of Chile) for [...] Read more.
Land surface temperature (LST) is one of the most important variables in the physical processes of surface energy and water balance. The temporal behavior of LST was analyzed between the latitudes 32°00′ S and 34°24′ S (Valparaíso and Metropolitana regions of Chile) for three summer months (December, January, and February) in the 2000–2017 period, using the Terra MODIS image information and applying the Mann–Kendall test. The results show an increase in LST in the study area, particularly in the Andes mountain range in January (5240 km2), which mainly comprises areas devoid of vegetation and eternal snow and glaciers, and are zones that act as water reserves for the capital city of Santiago. Similarly, vegetated areas such as forests, grasslands, and shrublands also show increasing trends in LST but over smaller surfaces. Because this study is regional, it is recommended to improve the spatial and temporal resolutions of the images to obtain conclusions on more local scales. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area. Source: Own elaboration.</p>
Full article ">Figure 2
<p>Methodological scheme of the study. Source: Own elaboration. (<a href="https://lpdaac.usgs.gov/dataset_discovery/modis" target="_blank">https://lpdaac.usgs.gov/dataset_discovery/modis</a>, accessed on 15 March 2024).</p>
Full article ">Figure 3
<p>Space–time distribution of areas with significant trends with a 95% confidence level in land surface temperatures (LSTs), based on the <span class="html-italic">p</span>-value of the Mann–Kendall test. (<b>1</b>): December; (<b>2</b>): January; (<b>3</b>): February. Source: Own elaboration from MODIS LST images.</p>
Full article ">Figure 4
<p>Spatial distribution of the linear regression analysis of the LST series for the month of December. (<b>1</b>): Trend values 2000–2016; (<b>2</b>): Positive and negative trends of the 2000–2016 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 5
<p>Spatial distribution of the linear regression analysis of the LST series for the month of January. (<b>1</b>): Trend values 2001–2017; (<b>2</b>): Positive and negative trends of the 2001–2017 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 6
<p>Spatial distribution of the linear regression analysis of the LST series for the month of February. (<b>1</b>): Trend values 2001–2017; (<b>2</b>): Positive and negative trends of the 2001–2017 time series, only in areas with significant trends and regression slope statistics. In degrees Celsius per year. Source: Own elaboration.</p>
Full article ">Figure 7
<p>(<b>1</b>): Land use for the entire study area; (<b>2</b>): Current land use in areas with significant trends for the month of December; (<b>3</b>): Current land use in areas with significant trends for the month of January; (<b>4</b>): Current land use in areas with significant trends for the month of February. Source: Own elaboration based on the study “Inventory of native vegetational resources of Chile, for Regions V, VI and RM” CONAF [<a href="#B38-hydrology-11-00103" class="html-bibr">38</a>].</p>
Full article ">Figure 8
<p>(<b>1</b>): Geomorphology of the study area; (<b>2</b>): Geomorphology in areas with significant trends for December; (<b>3</b>): Geomorphology in areas with significant trends for January; (<b>4</b>): Geomorphology in areas with significant trends for February. Source: Own elaboration based on the study “Geomorphological units of Chile” [<a href="#B48-hydrology-11-00103" class="html-bibr">48</a>].</p>
Full article ">Figure 9
<p>Behavior of the average LST by year, month, and land use for the analyzed period within the study area. The red line represents the linear trend.</p>
Full article ">
29 pages, 17646 KiB  
Article
Dust Events over the Urmia Lake Basin, NW Iran, in 2009–2022 and Their Potential Sources
by Abbas Ranjbar Saadat Abadi, Karim Abdukhakimovich Shukurov, Nasim Hossein Hamzeh, Dimitris G. Kaskaoutis, Christian Opp, Lyudmila Mihailovna Shukurova and Zahra Ghasabi
Remote Sens. 2024, 16(13), 2384; https://doi.org/10.3390/rs16132384 - 28 Jun 2024
Viewed by 690
Abstract
Nowadays, dried lake beds constitute the largest source of saline dust storms, with serious environmental and health issues in the surrounding areas. In this study, we examined the spatial–temporal distribution of monthly and annual dust events of varying intensity (dust in suspension, blowing [...] Read more.
Nowadays, dried lake beds constitute the largest source of saline dust storms, with serious environmental and health issues in the surrounding areas. In this study, we examined the spatial–temporal distribution of monthly and annual dust events of varying intensity (dust in suspension, blowing dust, dust storms) in the vicinity of the desiccated Urmia Lake in northwestern (NW) Iran, based on horizontal visibility data during 2009–2022. Dust in suspension, blowing dust and dust storm events exhibited different monthly patterns, with higher frequencies between March and October, especially in the southern and eastern parts of the Urmia Basin. Furthermore, the intra-annual variations in aerosol optical depth at 500 nm (AOD550) and Ångström exponent at 412/470 nm (AE) were investigated using Terra/Aqua MODIS (Moderate Resolution Imaging Spectroradiometer) data over the Urmia Lake Basin (36–39°N, 44–47°E). Monthly distributions of potential coarse aerosol (AE < 1) sources affecting the lower troposphere over the Urmia Basin were reconstructed, synergizing Terra/Aqua MODIS AOD550 for AE < 1 values and HYSPLIT_4 backward trajectories. The reconstructed monthly patterns of the potential sources were compared with the monthly spatial distribution of Terra MODIS AOD550 in the Middle East and Central Asia (20–70°E, 20–50°N). The results showed that deserts in the Middle East and the Aral–Caspian arid region (ACAR) mostly contribute to dust aerosol load over the Urmia Lake region, exhibiting higher frequency in spring and early summer. Local dust sources from dried lake beds further contribute to the dust AOD, especially in the western part of the Urmia Basin during March and April. The modeling (DREAM8-NMME-MACC) results revealed high concentrations of near-surface dust concentrations, which may have health effects on the local population, while distant sources from the Middle East are the main controlling factors to aerosol loading over the Urmia Basin. Full article
Show Figures

Figure 1

Figure 1
<p>The topography of the study area and locations of 26 weather stations in the Urmia Basin, NW Iran. Black circles represent the main synoptic stations (MS) with 8 observations per day, while open circles symbolize the intermediate synoptic stations (IS) with 5 observations per day.</p>
Full article ">Figure 2
<p>A snapshot of the Urmia Lake region (September 2003). The cells of a 1° × 1° size are shown and were used to extract data series of the Terra/Aqua MODIS AOD<sub>550</sub> and Ångström exponent from 2009 to 2022. Parts of the dried bottom covered with salt are seen at the east (in N-cell) and southeast (CNTR-cell) boundaries of the lake.</p>
Full article ">Figure 3
<p>Monthly number of dust events with horizontal visibilities less than or equal to 5 km for widespread suspended dust (code 06) during 2009–2022.</p>
Full article ">Figure 4
<p>Monthly number of dust events with horizontal visibilities less than or equal to 5 km for blowing dust and dust storms (BDSS, as defined in <a href="#remotesensing-16-02384-t001" class="html-table">Table 1</a>) during 2009–2022.</p>
Full article ">Figure 5
<p>Number of horizontal visibilities reduced due to severe (<b>a</b>,<b>d</b>), moderate (<b>b</b>,<b>e</b>) and weak (<b>c</b>,<b>f</b>) intensity of WSD (WW06) and BDSS (WW_Oth).</p>
Full article ">Figure 6
<p>The 80th, 90th and 95th percentiles, as well as the maximum DSC (<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="sans-serif">μ</mi> <mi mathvariant="normal">g</mi> </mrow> <mo>/</mo> <mrow> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msup> </mrow> </mrow> </mrow> </semantics></math>) values of the surface concentration of suspended particles, predicted by the DREAM8-NMME-MACC model for urban and rural points in the study area, based on the range of changes in the AQI for PM<sub>10</sub> (<a href="#remotesensing-16-02384-t003" class="html-table">Table 3</a>).</p>
Full article ">Figure 7
<p>Intra-annual variations in the monthly average Ångström exponent (lines) and AOD<sub>550</sub> values (bars) from Terra MODIS over grid cells around the Urmia Basin during 2009–2022. (<b>a</b>) NW-cell. (<b>b</b>) W-cell. (<b>c</b>) SW-cell. (<b>d</b>) S-cell. (<b>e</b>) SE-cell. (<b>f</b>) E-cell. (<b>g</b>) NE-cell. (<b>h</b>) N-cell (evaluative graph). (<b>i</b>). CNTR-cell (evaluative graph). Red and light pink (for N and CNTR cells) lines indicate the annual mean AOD<sub>550</sub> in 2009–2022.</p>
Full article ">Figure 8
<p>Intra-annual variations in the monthly average AOD<sub>550</sub> values for the cases of AE &lt; 1 (gray bars) and monthly total number of AE &lt; 1 cases (blue bars) in 2009–2022 by Terra MODIS data. (<b>a</b>) NW-cell. (<b>b</b>) W-cell. (<b>c</b>) SW-cell. (<b>d</b>) S-cell. (<b>e</b>) SE-cell. (<b>f</b>) E-cell. (<b>g</b>) NE-cell. (<b>h</b>) N-cell (evaluative graph). (<b>i</b>) CNTR-cell (evaluative graph). Red lines represent the mean annual AOD<sub>550</sub> due to coarse aerosols only.</p>
Full article ">Figure 9
<p>Monthly mean regional ABL contribution to AOD<sub>550</sub> at AE &lt; 1.0 (CWT-AOD<sub>550</sub>) according to Aqua and Terra MODIS data for the Urmia Lake region (marked with a black rectangle) in 2009–2022: The spatial resolution is 0.5° × 0.5°. Here and after, the Aral Sea is shown in its boundaries of the 1960s.</p>
Full article ">Figure 10
<p>The spatial distributions of monthly mean Terra MODIS Deep Blue AOD<sub>550</sub> during 2009–2022. The Urmia Lake region is marked with a black rectangle. The spatial resolution is 1° × 1°.</p>
Full article ">Figure A1
<p>Intra-annual variations in the monthly average Ångström exponent (lines) and the monthly average AOD<sub>550</sub> (bars) over the following cells in 2009–2021 by Aqua MODIS data. (<b>a</b>) NW-cell. (<b>b</b>) W-cell. (<b>c</b>) SW-cell. (<b>d</b>) S-cell. (<b>e</b>) SE-cell. (<b>f</b>) E-cell. (<b>g</b>) NE-cell. (<b>h</b>) N-cell (evaluative graph). (<b>i</b>) CNTR-cell (evaluative graph). Blue and red (light pink for N and CNTR cells) lines indicate, respectively, the annual mean Ångström exponent and AOD<sub>550</sub> values in 2009–2022.</p>
Full article ">Figure A2
<p>Intra-annual variations in the monthly average AOD<sub>550</sub> (gray bars) and monthly total number of coarse aerosol (Ångström exponent &lt; 1) events N (blue bars) in 2009–2022 by Aqua MODIS data. (<b>a</b>) NW-cell. (<b>b</b>) W-cell. (<b>c</b>) SW-cell. (<b>d</b>) S-cell. (<b>e</b>) SE-cell. (<b>f</b>) E-cell. (<b>g</b>) NE-cell. (<b>h</b>) N-cell (evaluative graph). (<b>i</b>) CNTR-cell (evaluative graph). Red (and light pink for N and CNTR cells) lines represent the mean annual AOD<sub>550</sub> averaged by coarse aerosol events.</p>
Full article ">Figure A3
<p>The spatial distributions of the monthly mean Terra-MODIS AOD<sub>550</sub> (Combined Deep Blue/Dark Target) during 2009–2022. The Urmia Lake region is marked with a black rectangle. The spatial resolution is 1° × 1°.</p>
Full article ">
31 pages, 8405 KiB  
Article
Lake and Atmospheric Heatwaves Caused by Extreme Dust Intrusion in Freshwater Lake Kinneret in the Eastern Mediterranean
by Pavel Kishcha, Yury Lechinsky and Boris Starobinets
Remote Sens. 2024, 16(13), 2314; https://doi.org/10.3390/rs16132314 - 25 Jun 2024
Viewed by 2294
Abstract
The role of dust intrusions in the formation of lake heatwaves has not yet been discussed in previous publications. We investigated a lake heatwave (LHW) and an atmospheric heatwave (AHW) in the freshwater Lake Kinneret in the Eastern Mediterranean: these were caused by [...] Read more.
The role of dust intrusions in the formation of lake heatwaves has not yet been discussed in previous publications. We investigated a lake heatwave (LHW) and an atmospheric heatwave (AHW) in the freshwater Lake Kinneret in the Eastern Mediterranean: these were caused by an extreme dust intrusion that lasted for a 10-day period (7–17 September 2015). The AHW and LHW were defined as periods of abnormally high air temperature (Tair) and lake surface water temperature (SWT) compared to their 90th percentile thresholds in September. In the daytime, the maximal intensities of AHW and LHW reached 3 °C and 2 °C, respectively. This was despite the pronounced drop in solar radiation due to the dust radiative effect. The satellite SWT retrievals were incapable of representing the abnormally high SWT in the presence of the extreme dust intrusion. Both METEOSAT and MODIS-Terra showed a sharp decrease in the SWT compared to the actual SWT: up to 10 °C in the daytime and up to 15 °C in the nighttime. Such a significant underestimation of the actual SWT in the presence of a dust intrusion should be considered when using satellite data to analyze heatwaves. In the absence of moisture advection, the AHW and LHW were accompanied by an increase of up to 30% in absolute humidity (ρv) over the lake. Being a powerful greenhouse gas, water vapor (characterized by an increased ρv) absorbed most of both the upwelling and downwelling longwave thermal radiation, heating the near-ground atmospheric layer (which is in direct contact with the lake water surface), in the daytime and nighttime. In the nighttime, the maximal intensity of the AHW and LHW reached 4 °C and 3 °C, respectively. Because of the observed steadily increasing dust pollution over the Eastern Mediterranean during the past several decades, we anticipate that dust-related lake heatwaves will intensify adverse effects on aquatic ecosystems such as reducing fishery resources and increasing harmful cyanobacteria blooms. Full article
(This article belongs to the Section Biogeosciences Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Map of Mediterranean SST trends (updated from Pisano et al. [<a href="#B8-remotesensing-16-02314" class="html-bibr">8</a>]). (<b>b</b>) Topography of the southeast Mediterranean region (31.8°N–33.6°N; 34.2°E–35.8°E) with Lake Kinneret, (<b>c</b>) bathymetric map of Lake Kinneret (−215 to −250 m a.s.l.). The green, blue, yellow, and purple pentagons designate the location of the following stations: Zemah (32.70°N, 35.58°E), Deir Hanna (32.86°N, 35.37°E), Ayyellet Hashahar (33.02°N, 35.57°E), and Avne Etan (32.81°N, 35.76°E), respectively. The black square shows the location of the Afula PM10 monitoring site (32.59°N, 35.27°E). The open circle shows the location of the Bet Dagan (32.00°N, 34.81°E) station. MH designates Mount Hermon, while A (32.82°N, 35.60°E, 40 m depth) designates the location of the monitoring station conducting measurements of water temperature and meteorological parameters in the lake. The blue rectangles designate two pixels on the 0.05° × 0.05° METEOSAT grid: they represent the water area in the lake (32.775°N–32.875°N; 35.575°E–35.625°E) where METEOSAT SWT was analyzed.</p>
Full article ">Figure 2
<p>Time series of 10 min PM10 concentration measurements taken at the Afula monitoring site: (<b>a</b>) during the period from 6 to 7 September and (<b>b</b>) during the period from 8 to 12 September 2015.</p>
Full article ">Figure 3
<p>Day-to-day variations in MODIS Deep Blue AOD averaged over North Israel (including Lake Kinneret) (32°N–33°N, 35°E–36°E) in September 2015. The absence of AOD measurements on 21 September 2015 can be explained by the presence of clouds over North Israel, which prevented MODIS from conducting AOD measurements.</p>
Full article ">Figure 4
<p>Comparison between day-to-day variations in maximal solar radiation (SR-MAX) in September 2015 and those of normal SR-MAX. The latter was defined as the average SR-MAX over the reference Septembers from 2011 to 2023 (excluding September 2015 and 2020). The short vertical lines designate the standard deviation of normal SR.</p>
Full article ">Figure 5
<p>Comparison between day-to-day variations in (<b>a</b>) DLWR-MAX and the 90th percentile threshold for DLWR-MAX; (<b>b</b>) ULWR-MAX and the 90th percentile threshold for ULWR-MAX; (<b>c</b>) DLWR-MIN and the 90th percentile threshold for DLWR-MIN; and (<b>d</b>) ULWR-MIN and the 90th percentile threshold for ULWR-MIN, based on in-situ radiometer measurements at St. A at Lake Kinneret in September 2015.</p>
Full article ">Figure 6
<p>Vertical temperature profiles from radiosonde soundings at St. Bet Dagan at 12 UTC on (<b>a</b>) 7 September, (<b>b</b>) 8 September, (<b>c</b>) 9 September, (<b>d</b>) 10 September, (<b>e</b>) 11 September, (<b>f</b>) 12 September, (<b>g</b>) 13 September, (<b>h</b>) 14 September, and (<b>i</b>) 15 September 2015.</p>
Full article ">Figure 7
<p>Comparison between day-to-day variations in daily maximal air temperature (Tair-MAX) in September 2015 and the 90th percentile threshold for Tair-MAX based on in-situ measurements at St. A in Lake Kinneret. On 3 September 2015, there were no in-situ Tair measurements in the daytime due to technical reasons.</p>
Full article ">Figure 8
<p>(<b>a</b>) Comparison between day-to-day variations in minimal air temperature (Tair-MIN) in September 2015 and the 90th percentile threshold for Tair-MIN, based on in-situ measurements at St. A in Lake Kinneret. (<b>b</b>) Comparison between day-to-day variations in Tair-MIN and its corresponding absolute humidity (ρ<sub>v</sub>) in September 2015 at St. A.</p>
Full article ">Figure 9
<p>Diurnal variations in the in-situ measured Lake Kinneret SWT from 6 to 11 September 2015.</p>
Full article ">Figure 10
<p>Comparison between day-to-day variations in (<b>a</b>) maximal Tair (Tair-MAX) and maximal SWT (SWT-MAX), and (<b>b</b>) minimal Tair (Tair-MIN) and minimal SWT (SWT-MIN) in September 2015. (<b>c</b>) Scatterplot between the daily temperature differences for every pair of consecutive days of Tair-MIN in September 2015 and those of SWT-MIN. <span class="html-italic">R2</span> designates the coefficient of determination, and <span class="html-italic">p</span> designates the significance level of the linear fit.</p>
Full article ">Figure 11
<p>(<b>a</b>) Comparison between day-to-day variations in in-situ SWT-MAX in September 2015 and the 90th percentile threshold for SWT-MAX, based on in-situ measurements at St. A on Lake Kinneret. (<b>b</b>) Comparison between day-to-day variations in in-situ SWT-MIN in September 2015 and the 90th percentile threshold for SWT-MIN, based on in-situ measurements at St. A.</p>
Full article ">Figure 12
<p>Day-to-day variations in in-situ SWT, satellite-based MODIS-Terra SWT, and METEOSAT SWT: (<b>a</b>) in the daytime (10 LT–11 LT) and (<b>b</b>) in the nighttime (22 LT–23 LT) in September 2015. The vertical lines designate the uncertainty of METEOSAT SWT.</p>
Full article ">Figure 13
<p>Comparison between day-to-day variations in (<b>a</b>) daily maximal SWT (SWT-MAX) and daily maximal WT-20 cm (WT-20cm-MAX), and (<b>b</b>) daily minimal SWT (SWT-MIN) and daily minimal WT-20 cm (WT-20cm-MIN), in September 2015.</p>
Full article ">Figure 14
<p>Comparison between day-to-day variations in (<b>a</b>) daily maximal water temperature at a depth of 20 cm (WT-20cm-MAX) and the 90th percentile threshold for WT-20cm-MAX, and (<b>b</b>) daily minimum water temperature at a depth of 20 cm (WT-20cm-MIN) and the 90th percentile threshold for WT-20cm-MIN, based on in-situ measurements in September 2015, at St. A on Lake Kinneret.</p>
Full article ">Figure 15
<p>Comparison between day-to-day variations in daily maximal absolute humidity (ρ<sub>v</sub>-MAX) and daily maximal mixing ratio (MIXR-MAX) in September 2015.</p>
Full article ">Figure 16
<p>(<b>a</b>) Comparison between day-to-day variations in daily maximal absolute humidity (ρ<sub>v</sub>-MAX) in September 2015 and 90th percentile for ρ<sub>v</sub>-MAX, based on in-situ measurements at St. A on Lake Kinneret. (<b>b</b>) Comparison between day-to-day variations in daily ρ<sub>v</sub>-MAX measured at St. A on Lake Kinneret and day-to-day variations in temperature inversion height (based on radiosonde soundings) in September 2015.</p>
Full article ">Figure 17
<p>Comparison between day-to-day variations in daily maximal absolute humidity (ρ<sub>v</sub>-MAX) in September 2015 and the 90th percentile threshold for ρ<sub>v</sub>-MAX at four meteorological stations: (<b>a</b>) St. A located on the lake (32.82°N, 35.60°E); (<b>b</b>) St. B represents the Ayyellet Hashahar station (33.02°N, 35.57°E); (<b>c</b>) St. C is the Deir Hanna station (32.86°N, 35.37°E); and (<b>d</b>) St. D is the Avne Etan station (32.81°N, 35.76°E). (<b>e</b>) Lake Kinneret region map with the location of meteorological stations.</p>
Full article ">Figure 18
<p>Correlation coefficients between SWT-MAX and ρ<sub>v</sub>-MAX during the period of 7 to 17 September in 2013, 2014, 2015, 2016, and 2017. The vertical lines designate the standard error of the correlation coefficient. One can detect the absence of a correlation between SWT-MAX and ρ<sub>v</sub>-MAX in September 2015.</p>
Full article ">Figure A1
<p>Comparison between day-to-day variations in Tair-MAX in September 2015 and the 90th percentile threshold for Tair-MAX, at four meteorological stations: (<b>a</b>) St. A located on the lake (32.82°N, 35.60°E); (<b>b</b>) St. B represents the Ayyellet Hashahar station (33.02°N, 35.57°E); (<b>c</b>) St. C is the Deir Hanna station (32.86°N, 35.37°E); and (<b>d</b>) St. D is the Avne Etan station (32.81°N, 35.76°E). (<b>e</b>) a map of the Lake Kinneret region with the location of meteorological stations.</p>
Full article ">Figure A2
<p>Comparison between day-to-day variations in Tair-MIN in September 2015 and the 90th percentile threshold for Tair-MIN at four meteorological stations: (<b>a</b>) St. A located in the lake (32.82°N, 35.60°E); (<b>b</b>) St. B represents the Ayyellet Hashahar station (33.02°N, 35.57°E); (<b>c</b>) St. C is the Deir Hanna station (32.86°N, 35.37°E); and (<b>d</b>) St. D is the Avne Etan station (32.81°N, 35.76°E). (<b>e</b>) a map of the Lake Kinneret region with the location of meteorological stations.</p>
Full article ">
18 pages, 4918 KiB  
Article
Assessment of Accuracy of Moderate-Resolution Imaging Spectroradiometer Sea Surface Temperature at High Latitudes Using Saildrone Data
by Chong Jia, Peter J. Minnett and Malgorzata Szczodrak
Remote Sens. 2024, 16(11), 2008; https://doi.org/10.3390/rs16112008 - 3 Jun 2024
Viewed by 1049
Abstract
The infrared (IR) satellite remote sensing of sea surface skin temperature (SSTskin) is challenging in the northern high-latitude region, especially in the Arctic because of its extreme environmental conditions, and thus the accuracy of SSTskin retrievals is questionable. Several Saildrone [...] Read more.
The infrared (IR) satellite remote sensing of sea surface skin temperature (SSTskin) is challenging in the northern high-latitude region, especially in the Arctic because of its extreme environmental conditions, and thus the accuracy of SSTskin retrievals is questionable. Several Saildrone uncrewed surface vehicles were deployed at the Pacific side of the Arctic in 2019, and two of them, SD-1036 and SD-1037, were equipped with a pair of IR pyrometers on the deck, whose measurements have been shown to be useful in the derivation of SSTskin with sufficient accuracy for scientific applications, providing an opportunity to validate satellite SSTskin retrievals. This study aims to assess the accuracy of MODIS-retrieved SSTskin from both Aqua and Terra satellites by comparisons with collocated Saildrone-derived SSTskin data. The mean difference in SSTskin from the SD-1036 and SD-1037 measurements is ~0.4 K, largely resulting from differences in the atmospheric conditions experienced by the two Saildrones. The performance of MODIS on Aqua and Terra in retrieving SSTskin is comparable. Negative brightness temperature (BT) differences between 11 μm and 12 μm channels are identified as being physically based, but are removed from the analyses as they present anomalous conditions for which the atmospheric correction algorithm is not suited. Overall, the MODIS SSTskin retrievals show negative mean biases, −0.234 K for Aqua and −0.295 K for Terra. The variations in the retrieval inaccuracies show an association with diurnal warming events in the upper ocean from long periods of sunlight in the Arctic. Also contributing to inaccuracies in the retrieval is the surface emissivity effect in BT differences characterized by the Emissivity-introduced BT difference (EΔBT) index. This study demonstrates the characteristics of MODIS-retrieved SSTskin in the Arctic, at least at the Pacific side, and underscores that more in situ SSTskin data at high latitudes are needed for further error identification and algorithm development of IR SSTskin. Full article
Show Figures

Figure 1

Figure 1
<p>The cruise trajectories of two NASA-funded Saildrone vehicles, SD-1036 (white) and SD-1037 (magenta), deployed during the 2019 Arctic Cruise from 15 May to 11 October. The background SST map is taken from the Multiscale Ultrahigh Resolution (MUR) Level-4 SST analysis data [<a href="#B37-remotesensing-16-02008" class="html-bibr">37</a>] on 16 September 2019. The subplot is a picture of the Saildrones at the starting point, which is courtesy of Saildrone Inc.</p>
Full article ">Figure 2
<p>The Aqua MODIS–Saildrone SST<sub>skin</sub> difference as a function of the (<b>a</b>) distance and (<b>b</b>) time difference in the allowed spatial–temporal intervals in the matchup criteria. Data include both SD-1036 and SD-1037 measurements. The one-to-one matchups were determined based on the smallest separation between the Saildrone measurement and MODIS pixel. The black linear fitted lines are given with the expression in the top right corner. (<b>c</b>,<b>d</b>) are similar to (<b>a</b>,<b>b</b>), but for the one-to-one matchups determined by the closest timestamp. The regressions in (<b>a</b>,<b>c</b>) pass the significance test at a 95% confidence level, but slopes in (<b>b</b>,<b>d</b>) are not significantly different from zero.</p>
Full article ">Figure 3
<p>(<b>a</b>) Histogram (normal distribution fitted curve in blue) of the Aqua MODIS–Saildrone SST<sub>skin</sub> difference and (<b>b</b>) the scatter plot of Saildrone- and MODIS-derived SST<sub>skin</sub> colored by the data density. (<b>c</b>,<b>d</b>) are similar to (<b>a</b>,<b>b</b>), but for Terra MODIS–Saildrone matchups.</p>
Full article ">Figure 4
<p>Histograms of the (<b>a</b>) Aqua MODIS BT difference between 11 μm and 12 μm channels and (<b>b</b>) air–sea temperature difference (ASTD) for the matchup data during the SD-1036 (light blue) and SD-1037 (light red) cruises. (<b>c</b>) The data density scatter plot of the BT difference and ASTD in Aqua MODIS–Saildrone matchups with the fitted dashed line.</p>
Full article ">Figure 5
<p>Maps of ASTD for Aqua MODIS–Saildrone matchups for (<b>a</b>) SD-1036 and (<b>b</b>) SD-1037. (<b>c</b>) The bivariate histogram for the longitude and latitude of the Aqua MODIS pixels matched with the SD-1036 (blue) and SD-1037 (red) measurements. The marks on some red columns indicate the heights of corresponding blue bars overwhelmed by the red ones.</p>
Full article ">Figure 6
<p>Reanalysis data from MERRA-2 matched with the Aqua MODIS–Saildrone matchups during the SD-1036 and SD-1037 cruises showing the vertical profiles of (<b>a</b>) specific humidity and (<b>b</b>) air temperature plotted as the mean (line and dots) ±1 standard deviation (envelope), as well as (<b>c</b>) the histogram of the total column water vapor.</p>
Full article ">Figure 7
<p>Histogram of RTTOV-simulated BT difference between 11 μm and 12 μm for Aqua MODIS pixels matched with Saildrone measurements during SD-1036 and SD-1037 cruises.</p>
Full article ">Figure 8
<p>Scatter plots (colored by data density) of the Aqua MODIS–Saildrone SST<sub>skin</sub> difference as a function of (<b>a</b>) the amplitude of diurnal warming with a fitted black dashed line when diurnal warming exists and (<b>b</b>) the Emissivity-introduced BT difference (EΔBT) with red dots and error bars indicating the mean and STD of temperature differences, calculated at 0.16 K intervals. The histogram distributions of diurnal warming and EΔBT are also plotted as the background for the data during SD-1036 and SD-1037 cruises separately.</p>
Full article ">
17 pages, 3972 KiB  
Article
Quantitative Assessment of Volcanic Thermal Activity from Space Using an Isolation Forest Machine Learning Algorithm
by Claudia Corradino, Arianna Beatrice Malaguti, Micheal S. Ramsey and Ciro Del Negro
Remote Sens. 2024, 16(11), 2001; https://doi.org/10.3390/rs16112001 - 1 Jun 2024
Cited by 1 | Viewed by 1174
Abstract
Understanding the dynamics of volcanic activity is crucial for volcano observatories in their efforts to forecast volcanic hazards. Satellite imager data hold promise in offering crucial insights into the thermal behavior of active volcanoes worldwide, facilitating the assessment of volcanic activity levels and [...] Read more.
Understanding the dynamics of volcanic activity is crucial for volcano observatories in their efforts to forecast volcanic hazards. Satellite imager data hold promise in offering crucial insights into the thermal behavior of active volcanoes worldwide, facilitating the assessment of volcanic activity levels and identifying significant changes during periods of volcano unrest. The Moderate Resolution Imaging Spectroradiometer (MODIS) sensor, aboard NASA’s Terra and Aqua satellites, provides invaluable data with high temporal and spectral resolution, enabling comprehensive thermal monitoring of eruptive activity. The accuracy of volcanic activity characterization depends on the quality of models used to relate the relationship between volcanic phenomena and target variables such as temperature. Under these circumstances, machine learning (ML) techniques such as decision trees can be employed to develop reliable models without necessarily offering any particular or explicit insights. Here, we present a ML approach for quantifying volcanic thermal activity levels in near real time using thermal infrared satellite data. We develop an unsupervised Isolation Forest machine learning algorithm, fully implemented in Google Colab using Google Earth Engine (GEE) which utilizes MODIS Land Surface Temperature (LST) data to automatically retrieve information on the thermal state of volcanoes. We evaluate the algorithm on various volcanoes worldwide characterized by different levels of volcanic activity. Full article
Show Figures

Figure 1

Figure 1
<p>Overview and representative image of the volcanoes studied. All images were captured via Google Earth Pro [<a href="http://www.earth.google.com" target="_blank">http://www.earth.google.com</a>; accessed on 27 March 2024] and QGIS [<a href="https://qgis.org/it/site/" target="_blank">https://qgis.org/it/site/</a>; accessed on 27 March 2024].</p>
Full article ">Figure 2
<p>Workflow of the proposed three-step approach.</p>
Full article ">Figure 3
<p>NLS adaptation phase to the anomaly detection phase for Etna. The reconstructed signal over the real thermal signal (<b>a</b>), the reconstructed error (<b>b</b>), the LST temperature of the detected anomalies (<b>c</b>), and the real thermal signal identified as anomalous over the real thermal signal (<b>d</b>).</p>
Full article ">Figure 4
<p>Time series of the temperature above average of the detected volcanic anomalies for Etna, Klyuchevskoy, and Lascar. Blue arrows indicate the starting of an eruption, while the gray shaded area indicates time windows characterized by an increase in activity preceding the eruption.</p>
Full article ">Figure 5
<p>Time series of the temperature above average of the detected volcanic anomalies for Fuego, Popocatépetl, and Stromboli. Blue arrows indicate the starting of an eruption, while the gray shaded area indicates time windows characterized by an increase in activity preceding the eruption.</p>
Full article ">Figure 6
<p>Time series of the temperature above average of the detected volcanic anomalies and the activity levels low (green bars), moderate (orange bars) and high (red bars) for Etna, Klyuchevskoy, Lascar, Fuego, Popocatépetl, and Stromboli.</p>
Full article ">
19 pages, 9247 KiB  
Article
Evaluating Visible–Infrared Imaging Radiometer Suite Imagery for Developing Near-Real-Time Nationwide Vegetation Cover Monitoring in Indonesia
by Yudi Setiawan, Kustiyo Kustiyo, Sahid Agustian Hudjimartsu, Judin Purwanto, Riva Rovani, Anna Tosiani, Ahmad Basyiruddin Usman, Tatik Kartika, Novie Indriasari, Lilik Budi Prasetyo and Belinda Arunarwati Margono
Remote Sens. 2024, 16(11), 1958; https://doi.org/10.3390/rs16111958 - 29 May 2024
Cited by 1 | Viewed by 834
Abstract
The necessity for precise and current data concerning the dynamics of land cover change in Indonesia is crucial for efforts to reduce natural vegetation cover due to agricultural expansion. The functionality of monitoring systems that incorporate Terra-MODIS is currently compromised by the limited [...] Read more.
The necessity for precise and current data concerning the dynamics of land cover change in Indonesia is crucial for efforts to reduce natural vegetation cover due to agricultural expansion. The functionality of monitoring systems that incorporate Terra-MODIS is currently compromised by the limited availability of data for the immediate future. This study seeks to assess the potential of VIIRS satellite imagery in developing an early warning system for monitoring vegetation cover change in Indonesia. The normalized differential open-area index (NDOAI) computed from 8-day VIIRS data was employed to detect changes in vegetation cover based on pixel-by-pixel subtraction in the NDOAI data time series. Evaluating the pixel-level accuracy of change detection is complicated due to the fact that we evaluate a change map at a coarser resolution than the Landsat-based reference map. The results revealed that increasing the threshold percentage is associated with improved accuracy. In change detection, there is often a trade-off between accuracy and sensitivity. A threshold that is too low may result in false positives, while a threshold that is too high may lead to missed changes. This study demonstrates that when a threshold value of less than 20% is applied, Landsat can identify vegetation cover changes at an earlier stage. Conversely, when a threshold value greater than 20% is employed, the VIIRS will detect the change 4.5 days earlier than Landsat. Additionally, the VIIRS is capable of detecting changes 25.4 days and 54.8 days faster than Landsat, respectively, when using thresholds of 40% and 70%. Full article
(This article belongs to the Special Issue Recent Progress in Remote Sensing of Land Cover Change)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) HDF tile coverage of VIIRS data for all of Indonesia; (<b>B</b>) 5 × 5-degree tile of Indonesia.</p>
Full article ">Figure 2
<p>Illustration of filtering procedure by (<b>A</b>) linear interpolation to estimate unknown values caused by clouds and (<b>B</b>) median moving window over time series datasets (Note: The blue line represents the original NDOI data, the dashed line indicates the filtered NDOAI data, and the green line shows the window size of the filter).</p>
Full article ">Figure 3
<p>Approach to a simple method for detecting vegetation changes based on differences in NDOAI values within a moving window every 8 days.</p>
Full article ">Figure 4
<p>Illustration of the change in the NDOAI value pattern from the previous year’s datasets and subsequent 8-day data to detect vegetation cover change every 8 days (Note: The red box indicates that the NDOAI is greater than the change threshold and has been marked as a change).</p>
Full article ">Figure 5
<p>(<b>A</b>) The 1 × 1-degree fishnet selected for developing data reference (cyan) and (<b>B</b>) VIIRS pixel size overlaid with Landsat ETM+ and a high-resolution image with a 100-point dot grid overlay.</p>
Full article ">Figure 6
<p>Illustrating the distribution of samples relative to the percentage of Landsat-based open areas.</p>
Full article ">Figure 7
<p>Distribution of vegetation cover change during 2022 detected by the devegetation method: (<b>A</b>) all regions of Indonesia, and selected sites in (<b>B</b>) Riau, (<b>C</b>) Jambi, (<b>D</b>) West Kalimantan and (<b>E</b>) East Kalimantan.</p>
Full article ">Figure 8
<p>The result of the devegetation analysis conducted in the year 2022 using (<b>A</b>) VIIRS and (<b>B</b>) Landsat.</p>
Full article ">Figure 9
<p>Comparison of devegetation analysis results based on (<b>A</b>) VIIRS, (<b>B</b>) Landsat, and (<b>C</b>) a combination of both for the selected site in Riau, Sumatra.</p>
Full article ">Figure 10
<p>A temporal comparison of the 16-day devegetation analysis results for the selected site in Riau, Sumatra, based on (<b>A</b>) VIIRS and (<b>B</b>) Landsat data.</p>
Full article ">Figure 11
<p>The disparity in detection speed between VIIRS and Landsat using different levels of thresholds as a reference point: (<b>A</b>) 5%, (<b>B</b>) 20%, (<b>C</b>) 40%, and (<b>D</b>) 70%.</p>
Full article ">Figure 12
<p>A comparison of different time frames between VIIRS and Landsat-based datasets, highlighting alterations in specific regions identifiable through the application of a 20% threshold.</p>
Full article ">Figure 13
<p>The relationship between accuracy, sample quantity, and the thresholds.</p>
Full article ">Figure 14
<p>Temporal discrepancy in detection time between VIIRS and Landsat based on the threshold level set by the percentage of open area within a 500 × 500 m grid size.</p>
Full article ">
25 pages, 8968 KiB  
Article
Consistency of Aerosol Optical Properties between MODIS Satellite Retrievals and AERONET over a 14-Year Period in Central–East Europe
by Lucia-Timea Deaconu, Alexandru Mereuță, Andrei Radovici, Horațiu Ioan Ștefănie, Camelia Botezan and Nicolae Ajtai
Remote Sens. 2024, 16(10), 1677; https://doi.org/10.3390/rs16101677 - 9 May 2024
Viewed by 1000
Abstract
Aerosols influence Earth’s climate by interacting with radiation and clouds. Remote sensing techniques aim to enhance our understanding of aerosol forcing using ground-based and satellite retrievals. Despite technological advancements, challenges persist in reducing uncertainties in satellite remote sensing. Our study examines retrieval biases [...] Read more.
Aerosols influence Earth’s climate by interacting with radiation and clouds. Remote sensing techniques aim to enhance our understanding of aerosol forcing using ground-based and satellite retrievals. Despite technological advancements, challenges persist in reducing uncertainties in satellite remote sensing. Our study examines retrieval biases in MODIS sensors on Terra and Aqua satellites compared to AERONET ground-based measurements. We assess their performance and the correlation with the AERONET aerosol optical depth (AOD) using 14 years of data (2010–2023) from 29 AERONET stations across 10 Central–East European countries. The results indicate discrepancies between MODIS Terra and Aqua retrievals: Terra overestimates the AOD at 16 AERONET stations, while Aqua underestimates the AOD at 21 stations. The examination of temporal biases in the AOD using the calculated estimated error (ER) between AERONET and MODIS retrievals reveals a notable seasonality in coincident retrievals. Both sensors show higher positive AOD biases against AERONET in spring and summer compared to fall and winter, with few ER values for Aqua indicating poor agreement with AERONET. Seasonal variations in correlation strength were noted, with significant improvements from winter to summer (from R2 of 0.58 in winter to R2 of 0.76 in summer for MODIS Terra and from R2 of 0.53 in winter to R2 of 0.74 in summer for MODIS Aqua). Over the fourteen-year period, monthly mean aerosol AOD trends indicate a decrease of −0.00027 from AERONET retrievals and negative monthly mean trends of the AOD from collocated MODIS Terra and Aqua retrievals of −0.00023 and −0.00025, respectively. An aerosol classification analysis showed that mixed aerosols comprised over 30% of the total aerosol composition, while polluted aerosols accounted for more than 22%, and continental aerosols contributed between 22% and 24%. The remaining 20% consists of biomass-burning, dust, and marine aerosols. Based on the aerosol classification method, we computed the bias between the AERONET AE and MODIS AE, which showed higher AE values for AERONET retrievals for a mixture of aerosols and biomass burning, while for marine aerosols, the MODIS AE was larger and for dust the results were inconclusive. Full article
(This article belongs to the Special Issue Remote Sensing of Aerosols, Planetary Boundary Layer, and Clouds)
Show Figures

Figure 1

Figure 1
<p>Spatial distribution of averaged aerosol optical depth (AOD) derived from MODIS Terra dataset over Central–East Europe for four seasons, (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn, between 2010 and 2023. The map also indicates the location of the 29 AERONET stations used for comparison. The size of the circles for the AERONET stations represents the number of data points available at each station, while the color bars indicate AOD values, cutoff at 0.25 for both MODIS and AERONET measurements.</p>
Full article ">Figure 2
<p>Spatial distribution of averaged AOD derived from MODIS Aqua dataset over the same region and time period as in <a href="#remotesensing-16-01677-f001" class="html-fig">Figure 1</a>, (<b>a</b>) winter, (<b>b</b>) spring, (<b>c</b>) summer, and (<b>d</b>) autumn. The figure layout, including the AERONET dataset and colour bar scale, remains consistent with <a href="#remotesensing-16-01677-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Comparison between the AERONET aerosol optical depth (AOD) at 550 nm and collocated MODIS Terra AOD at 550 nm for the four seasons (<b>a</b>) winter—DJF, (<b>b</b>) spring—MAM, (<b>c</b>) summer—JJA, and (<b>d</b>) autumn—SON, spanning from 2010 to 2023 over Central–East Europe. The scatter plots depict the statistical correlation between the two datasets for each season across all AERONET stations, with marginal histograms illustrating the distribution of AOD values for both datasets. The colour map indicates the corresponding AERONET Ångström exponent (AE) calculated with AOD at 440 and 870 nm. Linear equation values (slope and intercept), R values, R-square values, RMSE values, and number of points are shown in the table for each season.</p>
Full article ">Figure 4
<p>Same as <a href="#remotesensing-16-01677-f003" class="html-fig">Figure 3</a>, comparing the AERONET AOD at 550 nm with the collocated MODIS Aqua AOD at 550 nm, (<b>a</b>) winter—DJF, (<b>b</b>) spring—MAM, (<b>c</b>) summer—JJA, and (<b>d</b>) autumn—SON. The scatter plots, marginal histograms, colour map, and statistical metrics remain consistent with <a href="#remotesensing-16-01677-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 5
<p>Time series of monthly mean error ratios (ER in Equation (1)) (<b>a</b>) and the number of collocations (<b>b</b>) for the collocated dataset from 29 selected AERONET stations in CEE, time period 2010–2023. The Terra record is in red, and the Aqua is in blue. The number of collocations is season dependent (low number during winter months). Horizontal red lines are the ER value at −1 and 1.</p>
Full article ">Figure 6
<p>The distribution of AOD differences between AERONET and MODIS (<b>a</b>) Terra and (<b>b</b>) Aqua presented as probability density functions (PDF). Each of the 29 coloured plots represents the PDF of the ΔAOD (difference in daily AOD at 550 nm) between AERONET and MODIS retrievals for a specific location spanning from 2010 to 2023. The x-axis represents the ΔAOD values, ranging from negative to positive, while values on the colour bar represent the mean MODIS AOD bias at each individual location. The blue shades indicate a negative mean ΔAOD, while the red shades correspond to a positive mean ΔAOD. The vertical black line indicates the value 0 of ΔAOD.</p>
Full article ">Figure 7
<p>MODIS Terra (<b>a</b>) and Aqua (<b>b</b>) mean biases with confidence intervals (95%) specific for each AERONET location. Colour code same as <a href="#remotesensing-16-01677-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 8
<p>(<b>a</b>) The monthly mean aerosol optical depth (AOD) at 550 nm (blue line) with standard deviation and confidence interval at 95% for monthly means, spanning January 2010 to June 2023, alongside the monthly AOD trend (red line). (<b>b</b>) Similar for MODIS Terra and (<b>c</b>) MODIS Aqua. The multiannual mean AOD and standard deviation with confidence intervals are also shown.</p>
Full article ">Figure 9
<p>Relationship between Ångström exponent (AE) at wavelengths 440/870 nm and aerosol optical depth (AOD) at 440 nm. Six aerosol classes are distinguished based on thresholds of AE and AOD: biomass (green), continental (red), dust (orange), marine (blue), mixed (purple), and polluted (black).</p>
Full article ">Figure 10
<p>Pie charts illustrating aerosol type classification for total (<b>a</b>), urban (<b>b</b>), and rural (<b>c</b>) locations. Associated bar charts depict monthly aerosol type distribution spanning 2010–2023 across Central–East Europe for the three categories: total (<b>a</b>), urban (<b>b</b>), and rural (<b>c</b>). Aerosol types represented include marine (blue), dust (yellow), continental (red), mixed (purple), polluted (black), and biomass burning (green).</p>
Full article ">Figure 11
<p>Ångström exponent biases between AERONET and MODIS Terra retrievals for 6 aerosol types: biomass burning, continental, dust, marine, mixed, and polluted. The colour bar shows the points density. Descriptive statistics for these categories is also presented.</p>
Full article ">
18 pages, 3958 KiB  
Article
Drought Monitoring Using Moderate Resolution Imaging Spectroradiometer-Derived NDVI Anomalies in Northern Algeria from 2011 to 2022
by Ramzi Benhizia, Kwanele Phinzi, Fatemeh Hateffard, Haithem Aib and György Szabó
Environments 2024, 11(5), 95; https://doi.org/10.3390/environments11050095 - 4 May 2024
Viewed by 1695
Abstract
Drought has emerged as a major challenge to global food and water security, and is particularly pronounced for Algeria, which frequently grapples with water shortages. This paper sought to monitor and assess the temporal and spatial distribution of drought severity across northern Algeria [...] Read more.
Drought has emerged as a major challenge to global food and water security, and is particularly pronounced for Algeria, which frequently grapples with water shortages. This paper sought to monitor and assess the temporal and spatial distribution of drought severity across northern Algeria (excluding the Sahara) during the growing season from 2011 to 2022, while exploring the relationship between the normalized difference vegetation index (NDVI) anomaly and climate variables (rainfall and temperature). Temporal NDVI data from the Terra moderate resolution imaging spectroradiometer (MODIS) satellite covering the period 2000–2022 and climate data from the European Reanalysis 5th Generation (ERA5) datasets collected during the period 1990–2022 were used. The results showed that a considerable portion of northern Algeria has suffered from droughts of varying degrees of severity during the study period. The years 2022, 2021, 2016, and 2018 were the hardest hit, with 76%, 71%, 66%, and 60% of the area, respectively, experiencing drought conditions. While the relationship between the NDVI anomaly and the climatic factors showed variability across the different years, the steady decrease in vegetation health indicated by the NDVI anomaly corroborates the observed increase in drought intensity during the study period. We conclude that the MODIS-NDVI product offers a cost-efficient approach to monitor drought in data-scarce regions like Algeria, presenting a viable alternative to conventional climate-based drought indices, while serving as an initial step towards formulating drought mitigation plans. Full article
(This article belongs to the Special Issue Environmental Risk and Climate Change II)
Show Figures

Figure 1

Figure 1
<p>Location of the study area in northern Algeria (<b>a</b>). Land cover map of the northern part of Algeria and the location of weather stations falling within the study area (<b>b</b>).</p>
Full article ">Figure 2
<p>Spatial distribution of the mean NDVI for each year’s growing season from 2011 to 2022.</p>
Full article ">Figure 3
<p>Temporal variation in seasonal mean NDVI from 2011 to 2022 compared to the long-term NDVI (the red dashed line denotes a long-term median NDVI value).</p>
Full article ">Figure 4
<p>Percentage surface area of drought in northern Algeria from 2011 to 2022.</p>
Full article ">Figure 5
<p>The spatial distribution of drought in northern Algeria from 2011 to 2022.</p>
Full article ">Figure 6
<p>Seasonal variations in NDVI, rainfall, and temperature anomalies from 2011 to 2022.</p>
Full article ">Figure 7
<p>Spearman correlation coefficients (r<sub>s</sub>) of NDVI, rainfall, and temperature anomalies from 2011 to 2022.</p>
Full article ">
27 pages, 2083 KiB  
Article
A Wide-Angle Hyperspectral Top-of-Atmosphere Reflectance Model for the Libyan Desert
by Fuxiang Guo, Xiaobing Zheng, Yanna Zhang, Wei Wei, Zejie Zhang, Quan Zhang and Xin Li
Remote Sens. 2024, 16(8), 1406; https://doi.org/10.3390/rs16081406 - 16 Apr 2024
Viewed by 780
Abstract
Reference targets with stability, uniformity, and known reflectance on the Earth’s surface, such as deserts, can be used for the absolute radiometric calibration of satellite sensors. A wide-angle hyperspectral reflectance model at the top of atmosphere (TOA) over such a reference target will [...] Read more.
Reference targets with stability, uniformity, and known reflectance on the Earth’s surface, such as deserts, can be used for the absolute radiometric calibration of satellite sensors. A wide-angle hyperspectral reflectance model at the top of atmosphere (TOA) over such a reference target will expand the applicability of on-orbit calibration to different spectral bands and angles. To achieve the long-term, continuous, and high-precision absolute radiometric calibration of remote sensors, a wide-angle hyperspectral TOA reflectance model of the Libyan Desert was constructed based on spectral reflectance data, satellite overpass parameters, and atmospheric parameters from the Terra/Aqua and Earth Observation-1 (EO-1) satellites between 2003 and 2012. By means of angle fitting, viewing angle grouping, and spectral extension, the model is applicable for absolute radiometric calibration of the visible to short-wave infrared (SWIR) bands for sensors within viewing zenith angles of 65 degrees. To validate the accuracy and precision of the model, a total of 3120 long-term validations of model accuracy and 949 cross-validations with the Landsat 8 Operational Land Imager (OLI) and Suomi National Polar-Orbiting Partnership (NPP) Visible Infrared Imaging Radiometer Suite (VIIRS) satellite sensors between 2013 and 2020 were conducted. The results show that the TOA reflectance calculated by the model had a standard deviation (SD) of relative differences below 1.9% and a root-mean-square error (RMSE) below 0.8% when compared with observations from the Moderate Resolution Imaging Spectroradiometer (MODIS) and Landsat 8 OLI. The SD of the relative differences and the RMSE were within 2.7% when predicting VIIRS data. Full article
(This article belongs to the Special Issue Remote Sensing Satellites Calibration and Validation)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Libyan Desert in a satellite image. The overview map at the upper-right corner indicates the site’s location.</p>
Full article ">Figure 2
<p>Long-term clear-sky TOA reflectance in the red band of Landsat satellites.</p>
Full article ">Figure 3
<p>Long-term TOA reflectance of Aqua MODIS bands 1–7.</p>
Full article ">Figure 4
<p>Band coverage of the sensors.</p>
Full article ">Figure 5
<p>Daily nonuniformity distribution of the TOA reflectance and the iterative shrinking method.</p>
Full article ">Figure 6
<p>Distributions and sensitivity analyses of atmospheric parameters: AOD, water vapor, and ozone. (<b>a</b>) Distribution of AOD. (<b>b</b>) Relative difference analysis for AOD. (<b>c</b>) Distribution of water vapor. (<b>d</b>) Sensitivity analysis for water vapor. (<b>e</b>) Distribution of ozone. (<b>f</b>) Relative difference analysis for ozone. A positive relative difference corresponds to a brighter TOA reflectance.</p>
Full article ">Figure 7
<p>TOA reflectance as a function of the SZA and VZA in the blue band.</p>
Full article ">Figure 8
<p>Viewing angle distributions of Terra and Aqua.</p>
Full article ">Figure 9
<p>Hyperion hyperspectral TOA reflectance profiles.</p>
Full article ">Figure 10
<p>Spectral extension of the model results to Landsat 8 OLI bands <math display="inline"><semantics> <msubsup> <mi>ρ</mi> <mi>j</mi> <mrow> <mo>″</mo> </mrow> </msubsup> </semantics></math>.</p>
Full article ">Figure 11
<p>Spectral extension process.</p>
Full article ">Figure 12
<p>Observed and predicted TOA reflectance data for Terra MODIS.</p>
Full article ">Figure 13
<p>Relative differences between predictions and observations for Terra MODIS.</p>
Full article ">Figure 14
<p>Relative differences between predictions and observations for Aqua MODIS.</p>
Full article ">Figure 15
<p>Relative differences between predictions and observations for Landsat 8 OLI.</p>
Full article ">Figure 16
<p>Relative differences between predictions and observations for Suomi NPP VIIRS.</p>
Full article ">Figure 17
<p>Statistics of the model results. The positions of the circles along the y-axis represent the mean values of the relative differences. The error bars have a width of <math display="inline"><semantics> <mrow> <mo>±</mo> <mn>1</mn> </mrow> </semantics></math> SD around the circles.</p>
Full article ">
Back to TopTop