[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
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
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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,550)

Search Parameters:
Keywords = Sentinel-1A

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 14452 KiB  
Article
Detecting Melt Pond Onset on Landfast Arctic Sea Ice Using a Dual C-Band Satellite Approach
by Syeda Shahida Maknun, Torsten Geldsetzer, Vishnu Nandan, John Yackel and Mallik Mahmud
Remote Sens. 2024, 16(12), 2091; https://doi.org/10.3390/rs16122091 - 9 Jun 2024
Viewed by 750
Abstract
The presence of melt ponds on the surface of Arctic Sea ice affects its albedo, thermal properties, and overall melting rate; thus, the detection of melt pond onset is of significant importance for understanding the Arctic’s changing climate. This study investigates the utility [...] Read more.
The presence of melt ponds on the surface of Arctic Sea ice affects its albedo, thermal properties, and overall melting rate; thus, the detection of melt pond onset is of significant importance for understanding the Arctic’s changing climate. This study investigates the utility of a novel method for detecting the onset of melt ponds on sea ice using a satellite-based, dual-sensor C-band approach, whereby Sentinel-1 provides horizontally polarized (HH) data and Advanced SCATterometer (ASCAT) provides vertically polarized (VV) data. The co-polarized ratio (VV/HH) is used to detect the presence of melt ponds on landfast sea ice in the Canadian Arctic Archipelago in 2017 and 2018. ERA-5 air temperature and wind speed re-analysis datasets are used to establish the VV/HH threshold for pond onset detection, which have been further validated by Landsat-8 reflectance. The co-polarized ratio threshold of three standard deviations from the late winter season (April) mean co-pol ratio values are used for assessing pond onset detection associated with the air temperature and wind speed data, along with visual observations from Sentinel-1 and cloud-free Sentinel-2 imagery. In 2017, the pond onset detection rates were 70.59% for FYI and 92.3% for MYI. Results suggest that this method, because of its dual-platform application, has potential for providing large-area coverage estimation of the timing of sea ice melt pond onset using different earth observation satellites. Full article
Show Figures

Figure 1

Figure 1
<p>Maps of the study region showing sample site locations by ice type in (<b>a</b>) 2017 and (<b>b</b>) 2018. Sample site selection is described in <a href="#sec4dot3-remotesensing-16-02091" class="html-sec">Section 4.3</a>.</p>
Full article ">Figure 1 Cont.
<p>Maps of the study region showing sample site locations by ice type in (<b>a</b>) 2017 and (<b>b</b>) 2018. Sample site selection is described in <a href="#sec4dot3-remotesensing-16-02091" class="html-sec">Section 4.3</a>.</p>
Full article ">Figure 2
<p>ASCAT data processing method. (<b>a</b>) ASCAT 12.5 km points for one file: 1 April 2018 at 01:06. (<b>b</b>) ASCAT 12.5 km points for one day (01−04−2018). (<b>c</b>) ASCAT 12.5 km points following checks (purple dots) for one day (01−04−2018), a regular 5 km grid (black dots), and a 25 km land buffer (brown lines). (<b>d</b>) ASCAT daily weighted mean, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, for 2018-04-01. Legend is backscatter in dB.</p>
Full article ">Figure 3
<p>ERA−5 April mean daily 2 m air temperatures for all sites in 2017 and 2018.</p>
Full article ">Figure 4
<p>Wind speed data for (<b>a</b>) 2017 and (<b>b</b>) 2018 during the melt season.</p>
Full article ">Figure 5
<p>Site FYI_26_2017, showing daily time series of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, 2 m air temperature, and wind speed. The orange line in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math> data shows days with wind speed &lt; 3 ms<sup>−1</sup>. The navy−blue line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in sky blue. The black line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in dark grey. PO date (vertical red line) shows the date the pond onset occurred, in Day of Year. The validation window is presented as two separate yellow vertical lines.</p>
Full article ">Figure 6
<p>(<b>a</b>) Site FYI_26_2017 image from Sentinel-1 (Date: 2 April 2017). (<b>b</b>) Image from Sentinel-1 (Date: 25 May 2017). (<b>c</b>) Sentinel-2 RGB (4,3,2) image of the site from 12 June 2017. Blue circles in S-1 images and the red circle in S-2 represent the site locations.</p>
Full article ">Figure 7
<p>Site FYI_9_2018, showing daily time series of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, 2 m air temperature, and wind speed. The orange line in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math> data shows days with wind speed &lt; 3 ms<sup>−1</sup>. The navy−blue line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in sky blue. The black line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in dark grey. PO date (vertical red line) shows the date the pond onset occurred, in Day of Year. The validation window is presented as two separate yellow vertical lines.</p>
Full article ">Figure 8
<p>(<b>a</b>) Site FYI_9_2018 image from Sentinel-1 (Date: 3 April 2018). (<b>b</b>) Site image from Sentinel-1 (Date: 6 June 2018). (<b>c</b>) Sentinel-2 RGB (4,3,2) image of the site from 13 June 2018. Blue circles in S-1 images and the red circle in S-2 represent the site locations.</p>
Full article ">Figure 9
<p>Site MYI_11_2017, showing daily time series of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, 2 m air temperature, and wind speed. The orange line in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math> data shows days with wind speed &lt; 3 ms<sup>−1</sup>. The navy−blue line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in sky blue. The black line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in dark grey. PO date (vertical red line) shows the date the pond onset occurred, in Day of Year. The validation window is presented as two separate yellow vertical lines.</p>
Full article ">Figure 10
<p>(<b>a</b>) Site MYI_11_2017 image from Sentinel-1 (Date: 2 April 2017). (<b>b</b>) Image from Sentinel-1 (Date: 23 June 2017). (<b>c</b>) Sentinel-2 RGB (4,3,2) image of the site from 21 June 2017. Blue circles in S-1 images and the red circle in S-2 represent the site locations.</p>
Full article ">Figure 11
<p>Site MYI_13_2018, showing daily time series of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, 2 m air temperature, and wind speed. The orange line in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math> data shows days with wind speed &lt; 3 ms<sup>−1</sup>. The navy−blue line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in sky blue. The black line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in dark grey. PO date (vertical red line) shows the date the pond onset occurred, in Day of Year. The validation window is presented as two separate yellow vertical lines.</p>
Full article ">Figure 12
<p>(<b>a</b>) Site MYI_13_2018 image from Sentinel-1 (Date: 4 April 2018). (<b>b</b>) Image from Sentinel-1 (Date: 10 June 2018). (<b>c</b>) Sentinel-2 RGB (4,3,2) image of the site from 13 June 2018 (right). Blue circles in S-1 images and the red circle in S-2 represent the site locations.</p>
Full article ">Figure 13
<p>Box plots showing the regional and temporal variability in PO DOY for MYI and FYI. The black dots in the box plots represent observed PO on days that fell slightly outside the typical day range defined by the whiskers.</p>
Full article ">Figure 14
<p>Site FYI_04_2018, showing daily time series of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, 2 m air temperature, and wind speed. The orange line in the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>c</mi> <mi>o</mi> </mrow> </msub> </mrow> </semantics></math> data shows days with wind speed &lt; 3 ms<sup>−1</sup>. The navy−blue line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>H</mi> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in sky blue. The black line is <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>V</mi> <mi>V</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> </mrow> </semantics></math>, with imputed daily values in dark grey.</p>
Full article ">
19 pages, 13347 KiB  
Article
Quantitative Analysis of Spatiotemporal Patterns and Factor Contributions of Surface Ozone in the North China Plain
by Yi Li, Mengjiao Liu, Lingyue Lv, Jinhui Liang, Mingliang Ma, Mengnan Liu and Pingjie Fu
Appl. Sci. 2024, 14(12), 5026; https://doi.org/10.3390/app14125026 - 9 Jun 2024
Viewed by 637
Abstract
Over the past decade, surface ozone has emerged as a significant air pollutant in China, especially in the North China Plain (NCP). For effective ozone management in the NCP, it is crucial to accurately estimate the surface ozone levels and identify the primary [...] Read more.
Over the past decade, surface ozone has emerged as a significant air pollutant in China, especially in the North China Plain (NCP). For effective ozone management in the NCP, it is crucial to accurately estimate the surface ozone levels and identify the primary influencing factors for ozone pollution in this region. This study utilized ozone precursors such as volatile organic compounds (VOCs) and nitrogen oxides (NOX), meteorological data, land cover, normalized difference vegetation index (NDVI), terrain, and population data to build an extreme gradient boosting (XGBoost)-based ozone estimation model in the NCP during 2019 to 2021. Four ozone estimation models were developed using different NO2 and formaldehyde (HCHO) datasets from the Sentinel-5 TROPOMI observations and Copernicus Atmosphere Monitoring Service (CAMS) reanalysis data. Site-based validation results of these four models showed high accuracy with R2 values above 0.86. Among these four models, two models with higher accuracy and higher spatial coverage ratio were selected, and their results were averaged to produce the final ozone estimation products. The results indicated that VOCs and NOX were the two main pollutants causing ozone pollution in the NCP, and their relative contributions accounted for more than 23.34% and 10.23%, respectively, while HCHO also played a significant role, contributing over 5.64%. Additionally, meteorological factors also had a notable impact, contributing 28.63% to ozone pollution, with each individual factor contributing more than 2.38%. The spatial distribution of ozone pollution identified the Hebei–Shandong–Henan junction as a pollution hotspot, with the peak occurring in summer, particularly in June. Therefore, for this hotspot region in the NCP, promoting the reduction in VOCs and NOx can play an important role in the mitigation of O3 pollution and the improvement in air quality in this region. Full article
(This article belongs to the Special Issue Air Pollution and Its Impact on the Atmospheric Environment)
Show Figures

Figure 1

Figure 1
<p>Distribution of 133 ground-based observation sites in the NCP from 2019 to 2021. Different colors represent the foundation sites of different provinces, among them, red represents Shandong Province, green represents Hebei Province, blue represents Shanxi Province, pink represents Hebei Province, orange represents Tianjin City, cyan represents Beijing City, and yellow represents the Inner Mongolia Autonomous Region.</p>
Full article ">Figure 2
<p>For the site-based validation of the four MDA8 estimation models. (<b>a</b>) displays a comparative analysis at a selection of sites (representing 5% of the total), where the observed surface O<sub>3</sub> levels are juxtaposed with the O<sub>3</sub> estimates that were generated using S5P NO<sub>2</sub> and S5P HCHO data. (<b>b</b>) displays the validation findings for the model that incorporates S5P NO<sub>2</sub> and CAMS HCHO. (<b>c</b>) illustrates the results for the model utilizing CAMS NO<sub>2</sub> and S5P HCHO, and finally, (<b>d</b>) provides the validation data for the model that employs both CAMS NO<sub>2</sub> and CAMS HCHO. Each figure elucidates the performance of the respective model in predicting the surface O<sub>3</sub> concentrations compared to the ground-based measurements. The red dotted line represents the fitting results.</p>
Full article ">Figure 3
<p>Comparison of the relative importance of explanatory variables to the surface ozone estimation model in the NCP. (<b>a</b>) denotes the results derived from the surface ozone estimation model using S5P NO<sub>2</sub> and CAMS HCHO while (<b>b</b>) indicates the results for the model using CAMS NO<sub>2</sub> and CAMS HCHO.</p>
Full article ">Figure 4
<p>Spatial distribution of the seasonal averaged and annual averaged MDA8 concentration in the 2019–2021 period. (<b>a</b>–<b>d</b>) indicate the MDA8 in spring, summer, autumn, and winter, respectively. (<b>e</b>–<b>g</b>) indicate the annual averaged MDA8 in 2019, 2020, and 2021.</p>
Full article ">Figure 5
<p>Spatial patterns of MDA8 exceedances (MDA8 &gt; 160 µg m<sup>−3</sup>) and the month with the largest MDA8 exceedances in the 2019–2021 period. (<b>a</b>–<b>c</b>) indicate the MDA8 exceedances in 2019, 2020, and 2021. (<b>d</b>–<b>f</b>) represent the spatial distribution of the months with the highest MDA8 exceedances in 2019, 2020, and 2021.</p>
Full article ">Figure 6
<p>Temporal variations in the MDA8 exceedances ratio in each month and the accumulated MDA8 exceedances in each province from 2019 to 2021. (<b>a</b>) indicates the MDA8 exceedances ratio in each month in 2019, 2020, and 2021. (<b>b</b>) represents the accumulated MDA8 exceedances in each province in 2019, 2020, and 2021.</p>
Full article ">
25 pages, 35929 KiB  
Article
Identifying Plausible Labels from Noisy Training Data for a Land Use and Land Cover Classification Application in Amazônia Legal
by Maximilian Hell and Melanie Brandmeier
Remote Sens. 2024, 16(12), 2080; https://doi.org/10.3390/rs16122080 - 8 Jun 2024
Viewed by 469
Abstract
Most studies in the field of land use and land cover (LULC) classification in remote sensing rely on supervised classification, which requires a substantial amount of accurate label data. However, reliable data are often not immediately available, and are obtained through time-consuming manual [...] Read more.
Most studies in the field of land use and land cover (LULC) classification in remote sensing rely on supervised classification, which requires a substantial amount of accurate label data. However, reliable data are often not immediately available, and are obtained through time-consuming manual labor. One potential solution to this problem is the use of already available classification maps, which may not be the true ground truth and may contain noise from multiple possible sources. This is also true for the classification maps of the MapBiomas project, which provides land use and land cover (LULC) maps on a yearly basis, classifying the Amazon basin into more than 24 classes based on the Landsat data. In this study, we utilize the Sentinel-2 data with a higher spatial resolution in conjunction with the MapBiomas maps to evaluate a proposed noise removal method and to improve classification results. We introduce a novel noise detection method that relies on identifying anchor points in feature space through clustering with self-organizing maps (SOM). The pixel label is relabeled using nearest neighbor rules, or can be removed if it is unknown. A challenge in this approach is the quantification of noise in such a real-world dataset. To overcome this problem, highly reliable validation sets were manually created for quantitative performance assessment. The results demonstrate a significant increase in overall accuracy compared to MapBiomas labels, from 79.85% to 89.65%. Additionally, we trained the L2HNet using both MapBiomas labels and the filtered labels from our approach. The overall accuracy for this model reached 93.75% with the filtered labels, compared to the baseline of 74.31%. This highlights the significance of noise detection and filtering in remote sensing, and emphasizes the need for further research in this area. Full article
(This article belongs to the Section AI Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Study area (blue), spanning the UTM tiles 21LXE and 21LXF, within the bounds of the Amazon basin, or Amazônia Legal (green).</p>
Full article ">Figure 2
<p><b>Left</b>: True color rendering of used Sentinel-2 scene with minimal cloud cover. Extent: UTM tile 21LXE and 21LXF. It covers an area of 109.80 km × 209.82 km. <b>Right</b>: MapBiomas Collection 6 label map.</p>
Full article ">Figure 3
<p>Overview of the proposed method’s workflow. The input data and corresponding labels are flattened to tabular data, where each row represents a single pixel and the columns represent the bands. Those data are then split by their prior labels given by the label map. For each class, a separate SOM is learned to create anchor points for that class. The anchor points are then aggregated. For each data point, the <span class="html-italic">k</span> nearest anchor points in the feature space are searched and weighted. The label of the data point can then be kept, changed (if confident enough), or marked as unknown. The filtered and corrected labels can then be reconstructed as an image.</p>
Full article ">Figure 4
<p>Example of a contingency matrix for calculation of McNemar’s test.</p>
Full article ">Figure 5
<p>Confusion matrix using the validation data as ground truth compared to the MapBiomas label map Collection 6.</p>
Full article ">Figure 6
<p>Confusion matrix using the validation data as ground truth compared to the corrected label map created with the proposed approach. <span class="html-italic">k</span> was set to 10, and the neighbors were weighted by distance.</p>
Full article ">Figure 7
<p>Confusion matrix using the validation data as ground truth compared to the classification with a Random Forest classifier.</p>
Full article ">Figure 8
<p>Confusion matrix using the validation data as ground truth compared to the classification with the L2HNet.</p>
Full article ">Figure 9
<p>Confusion matrix using the validation data as ground truth compared to the classification of the L2HNet with the filtered data as training input.</p>
Full article ">Figure 10
<p>Qualitative assessment of selected areas showing in the columns: true color image of the used Sentinel-2 imagery, the MapBiomas map, the classification map corrected with the presented approach, the classification with the RF classifier, and the classification with the L2HNet. The rows reflect the five chosen locations for the qualitative assessment. It should be noted that each location is depicted at a different scale. The rows (<b>a</b>–<b>e</b>) correspond to the selected areas for qualitative validation marked in <a href="#remotesensing-16-02080-f0A1" class="html-fig">Figure A1</a>.</p>
Full article ">Figure 11
<p>Qualitative assessment of selected areas showing in the columns: true color image of the used Sentinel-2 imagery, the MapBiomas map, the classification map of the L2HNet trained with MapBiomas data, the classification with L2HNet using the filtered labels. The rows reflect the five chosen locations for the qualitative assessment. It should be noted that each location is depicted at a different scale. The rows (<b>a</b>–<b>e</b>) correspond to the selected areas for qualitative validation marked in <a href="#remotesensing-16-02080-f0A1" class="html-fig">Figure A1</a>.</p>
Full article ">Figure A1
<p>Spatial distribution of the manually created validation areas. Extent: UTM tile 21LXE and 21LXF. Area covered: 109.80 km × 209.82 km. The white bounding boxes a–e are showing the areas for qualitative validation (cf. <a href="#sec3dot2-remotesensing-16-02080" class="html-sec">Section 3.2</a>).</p>
Full article ">
21 pages, 4644 KiB  
Article
Super-Resolution Image Reconstruction Method between Sentinel-2 and Gaofen-2 Based on Cascaded Generative Adversarial Networks
by Xinyu Wang, Zurui Ao, Runhao Li, Yingchun Fu, Yufei Xue and Yunxin Ge
Appl. Sci. 2024, 14(12), 5013; https://doi.org/10.3390/app14125013 - 8 Jun 2024
Viewed by 626
Abstract
Due to the multi-scale and spectral features of remote sensing images compared to natural images, there are significant challenges in super-resolution reconstruction (SR) tasks. Networks trained on simulated data often exhibit poor reconstruction performance on real low-resolution (LR) images. Additionally, compared to natural [...] Read more.
Due to the multi-scale and spectral features of remote sensing images compared to natural images, there are significant challenges in super-resolution reconstruction (SR) tasks. Networks trained on simulated data often exhibit poor reconstruction performance on real low-resolution (LR) images. Additionally, compared to natural images, remote sensing imagery involves fewer high-frequency components in network construction. To address the above issues, we introduce a new high–low-resolution dataset GF_Sen based on GaoFen-2 and Sentinel-2 images and propose a cascaded network CSWGAN combined with spatial–frequency features. Firstly, based on the proposed self-attention GAN (SGAN) and wavelet-based GAN (WGAN) in this study, the CSWGAN combines the strengths of both networks. It not only models long-range dependencies and better utilizes global feature information, but also extracts frequency content differences between different images, enhancing the learning of high-frequency information. Experiments have shown that the networks trained based on the GF_Sen can achieve better performance than those trained on simulated data. The reconstructed images from the CSWGAN demonstrate improvements in the PSNR and SSIM by 4.375 and 4.877, respectively, compared to the relatively optimal performance of the ESRGAN. The CSWGAN can reflect the reconstruction advantages of a high-frequency scene and provides a working foundation for fine-scale applications in remote sensing. Full article
(This article belongs to the Section Earth Sciences)
Show Figures

Figure 1

Figure 1
<p>The self-attention mechanism for CSWGAN.</p>
Full article ">Figure 2
<p>Network architecture of CSWGAN.</p>
Full article ">Figure 3
<p>LR-HR image-pair samples in GF_Sen dataset. The first row is HR images from GaoFen-2, while the second row is the corresponding LR images from Sentinel-2.</p>
Full article ">Figure 4
<p>Spatial distribution of Sentinel-2 and GaoFen-2.</p>
Full article ">Figure 5
<p>Reconstruction results of different methods.</p>
Full article ">Figure 5 Cont.
<p>Reconstruction results of different methods.</p>
Full article ">Figure 6
<p>Reconstruction results of GF_Sen.</p>
Full article ">Figure 7
<p>Reconstruction results of USC-SIPI.</p>
Full article ">Figure 8
<p>Sample images used in frequency-domain feature statistics.</p>
Full article ">Figure 9
<p>Evaluation histogram of high-frequency characteristic images with different reconstruction results. (<b>a</b>) and (<b>b</b>), (<b>c</b>) and (<b>d</b>), (<b>e</b>) and (<b>f</b>) are the results of pictures 1–3, respectively.</p>
Full article ">
23 pages, 6314 KiB  
Article
Modeling Gross Primary Production (GPP) of a Mediterranean Grassland in Central Spain Using Sentinel-2 NDVI and Meteorological Field Information
by Víctor Cicuéndez, Rosa Inclán, Enrique P. Sánchez-Cañete, Carlos Román-Cascón, César Sáenz and Carlos Yagüe
Agronomy 2024, 14(6), 1243; https://doi.org/10.3390/agronomy14061243 - 7 Jun 2024
Viewed by 408
Abstract
Mediterranean grasslands provide different ecosystems and social and economic services to the Mediterranean basin. Specifically, in Spain, pastures occupy more than 55% of the Spanish surface. Farmers and policymakers need to estimate the Gross Primary Production (GPP) to make sustainable management of these [...] Read more.
Mediterranean grasslands provide different ecosystems and social and economic services to the Mediterranean basin. Specifically, in Spain, pastures occupy more than 55% of the Spanish surface. Farmers and policymakers need to estimate the Gross Primary Production (GPP) to make sustainable management of these ecosystems and to study the role of grasslands acting as sinks or sources of Carbon in the context of climate change. High-frequency satellites, such as Sentinel-2, have opened the door to study GPP with a higher spatial and lower revisit time (10 m and 5 days). Therefore, the overall objective of this research is to estimate an ecosystem light use efficiency (eLUE) GPP model for a Mediterranean grassland in central Spain using Sentinel-2 NDVI Normalized Difference Vegetation Index (NDVI), complemented with meteorological information at the field scale for a relatively long period (from January 2018 to July 2020). The GPP models studied in this research were the MODIS GPP product, as well as the four eLUE models built with MODIS or Sentinel-2 NDVI and complemented by the inclusion of minimum temperature (Tmin) and soil water content (SWC). The models were validated through the GPP obtained from an eddy-covariance flux tower located in the study site (GPP_T). Results showed that the MODIS GPP product underestimated the GPP_T of the grassland ecosystem. Besides this, the approach of the eLUE concept was valid for estimating GPP in this Mediterranean grassland ecosystem. In addition, the models showed an improvement using Sentinel-2 NDVI compared to MODIS GPP product and compared to the models that used MODIS NDVI due to its higher spatial and temporal resolution. The inclusion of Tmin and SWC was also a determinant in improving GPP models during winter and summer periods. This work also illustrates how the main wind directions of the study area must be considered to appropriately estimate the footprint of the eddy covariance flux tower. In conclusion, this study is the first step to efficiently estimating the GPP of Mediterranean grasslands using the Sentinel-2 NDVI with complementary meteorological field information to make the management of these ecosystems sustainable. Full article
(This article belongs to the Special Issue Advances in Grassland Productivity and Sustainability — Volume II)
Show Figures

Figure 1

Figure 1
<p>Location of the eddy covariance flux tower in Spain (<b>a</b>), location (yellow star) of the flux tower in the village of El Escorial in Madrid (central Spain) at 40°34′56″ N and 4°08′16″ W (<b>b</b>) (images from © Google Earth (Landsat/Copernicus)), the eddy covariance flux tower (<b>c</b>) and scheme of equipment of flux tower (<b>d</b>).</p>
Full article ">Figure 2
<p>Time series of tower Gross Primary Production (GPP_T) and Soil Water Content at 4 cm depth (SWC) from January 2018 to July 2020 in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 3
<p>Time series obtained for MODIS Gross Primary Production product (MODIS_GPP) from January 2018 to July 2020 versus the tower Gross Primary Production (GPP_T) in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 4
<p>Relationship between eLUE (GPP_T/PAR<sub>TOA</sub>) and MODIS NDVI from January 2018 to December 2019 in the Mediterranean grassland of La Herrería (<span class="html-italic">n</span> = 92, R<sup>2</sup> = 0.347, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">Figure 5
<p>Time series obtained for the Gross Primary Production MODIS model (GPP_M1) from January 2018 to July 2020 versus the tower Gross Primary Production (GPP_T) in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 6
<p>Values for the Gross Primary Production MODIS model 1 (GPP_M1) versus the tower Gross Primary Production (GPP_T) for the year 2020 (January to July) in the Mediterranean grassland of La Herrería (<span class="html-italic">n</span> = 25, R<sup>2</sup> = 0.78, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">Figure 7
<p>Time series obtained for the Gross Primary Production MODIS multiple linear regression model (GPP_M2) from January 2018 to July 2020 versus the tower Gross Primary Production (GPP_T) in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 8
<p>Values for model the Gross Primary Production MODIS multiple linear regression model (GPP_M2) versus the tower Gross Primary Production (GPP_T) for the year 2020 (January to July) in the Mediterranean grassland of La Herrería (<span class="html-italic">n</span> = 25, R<sup>2</sup> = 0.75, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">Figure 9
<p>Histogram of wind directions at 11.00 UTC from January 2018 to July 2020 in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 10
<p>The wind rose for the period using the wind values at 11.00 UTC from January 2018 to July 2020 in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 11
<p>The pixels selected for calculating the spatial average of NDVI (inside the green area, which was delimited by the main wind direction in the Mediterranean grassland of La Herrería. A circle of 110 m was drawn centered in the tower.</p>
Full article ">Figure 12
<p>Relationship between eLUE (GPP/PAR<sub>TOA</sub>) and Sentinel-2 NDVI from January 2018 to December 2019 in the Mediterranean grassland (<span class="html-italic">n</span> = 146, R<sup>2</sup> = 0.51, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">Figure 13
<p>Time series obtained for the Gross Primary Production Sentinel model (GPP_S1) from January 2018 to July 2020 versus the tower Gross Primary Production (GPP_T) in the Mediterranean grassland of La Herrería.</p>
Full article ">Figure 14
<p>Values for the Gross Primary Production Sentinel-2 model 1 (GPP_S1) versus the tower Gross Primary Production (GPP_T) for the year 2020 (January to July) in the Mediterranean grassland of La Herrería (<span class="html-italic">n</span> = 40, R<sup>2</sup> = 0.87, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">Figure 15
<p>Time series obtained for the Gross Primary Production Sentinel-2 multiple linear regression model (GPP_S2) from January 2018 to July 2020 versus the tower Gross Primary Production (GPP_T) in a Mediterranean grassland.</p>
Full article ">Figure 16
<p>Values for the Gross Primary Production Sentinel-2 multiple linear regression model (GPP_S2) versus the tower Gross Primary Production (GPP_T) for the year 2020 (January to July) in the Mediterranean grassland of La Herrería (<span class="html-italic">n</span> = 25, R<sup>2</sup> = 0.87, ANOVA <span class="html-italic">p</span>-value = 0.00).</p>
Full article ">
12 pages, 1749 KiB  
Systematic Review
Wire-Free Targeted Axillary Dissection: A Pooled Analysis of 1300+ Cases Post-Neoadjuvant Systemic Therapy in Node-Positive Early Breast Cancer
by Jajini Varghese, Neill Patani, Umar Wazir, Shonnelly Novintan, Michael J. Michell, Anmol Malhotra, Kinan Mokbel and Kefah Mokbel
Cancers 2024, 16(12), 2172; https://doi.org/10.3390/cancers16122172 - 7 Jun 2024
Viewed by 820
Abstract
Recent advances in neoadjuvant systemic therapy (NST) have significantly improved pathologic complete response rates in early breast cancer, challenging the role of axillary lymph node dissection in nose-positive patients. Targeted axillary dissection (TAD) integrates marked lymph node biopsy (MLNB) and tracer-guided sentinel lymph [...] Read more.
Recent advances in neoadjuvant systemic therapy (NST) have significantly improved pathologic complete response rates in early breast cancer, challenging the role of axillary lymph node dissection in nose-positive patients. Targeted axillary dissection (TAD) integrates marked lymph node biopsy (MLNB) and tracer-guided sentinel lymph node biopsy (SLNB). The introduction of new wire-free localisation markers (LMs) has streamlined TAD and increased its adoption. The primary endpoints include the successful localisation and retrieval rates of LMs. The secondary endpoints include the pathological complete response (pCR), SLNB, and MLNB concordance, as well as false-negative rates. Seventeen studies encompassing 1358 TAD procedures in 1355 met the inclusion criteria. The localisation and retrieval rate of LMs were 97% and 99%. A concordance rate of 67% (95% CI: 64–70) between SLNB and MLNB was demonstrated. Notably, 49 days (range: 0–272) was the average LM deployment time to surgery. pCR was observed in 46% (95% CI: 43–49) of cases, with no significant procedure-related complications. Omitting MLNB or SLNB would have under-staged the axilla in 15.2% or 5.4% (p = 0.0001) of cases, respectively. MLNB inclusion in axillary staging post-NST for initially node-positive patients is crucial. The radiation-free Savi Scout, with its minimal MRI artefacts, is the preferred technology for TAD. Full article
(This article belongs to the Special Issue Neo-Adjuvant Treatment of Breast Cancer)
Show Figures

Figure 1

Figure 1
<p>Successful localisation of localisation markers [<a href="#B13-cancers-16-02172" class="html-bibr">13</a>,<a href="#B14-cancers-16-02172" class="html-bibr">14</a>,<a href="#B15-cancers-16-02172" class="html-bibr">15</a>,<a href="#B16-cancers-16-02172" class="html-bibr">16</a>,<a href="#B17-cancers-16-02172" class="html-bibr">17</a>,<a href="#B18-cancers-16-02172" class="html-bibr">18</a>,<a href="#B19-cancers-16-02172" class="html-bibr">19</a>,<a href="#B20-cancers-16-02172" class="html-bibr">20</a>,<a href="#B21-cancers-16-02172" class="html-bibr">21</a>,<a href="#B22-cancers-16-02172" class="html-bibr">22</a>,<a href="#B23-cancers-16-02172" class="html-bibr">23</a>,<a href="#B24-cancers-16-02172" class="html-bibr">24</a>,<a href="#B25-cancers-16-02172" class="html-bibr">25</a>,<a href="#B26-cancers-16-02172" class="html-bibr">26</a>,<a href="#B27-cancers-16-02172" class="html-bibr">27</a>,<a href="#B28-cancers-16-02172" class="html-bibr">28</a>,<a href="#B29-cancers-16-02172" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Successful retrieval of localisation markers [<a href="#B13-cancers-16-02172" class="html-bibr">13</a>,<a href="#B14-cancers-16-02172" class="html-bibr">14</a>,<a href="#B15-cancers-16-02172" class="html-bibr">15</a>,<a href="#B16-cancers-16-02172" class="html-bibr">16</a>,<a href="#B17-cancers-16-02172" class="html-bibr">17</a>,<a href="#B19-cancers-16-02172" class="html-bibr">19</a>,<a href="#B20-cancers-16-02172" class="html-bibr">20</a>,<a href="#B21-cancers-16-02172" class="html-bibr">21</a>,<a href="#B22-cancers-16-02172" class="html-bibr">22</a>,<a href="#B23-cancers-16-02172" class="html-bibr">23</a>,<a href="#B24-cancers-16-02172" class="html-bibr">24</a>,<a href="#B25-cancers-16-02172" class="html-bibr">25</a>,<a href="#B26-cancers-16-02172" class="html-bibr">26</a>,<a href="#B27-cancers-16-02172" class="html-bibr">27</a>,<a href="#B29-cancers-16-02172" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>Pathological complete response [<a href="#B13-cancers-16-02172" class="html-bibr">13</a>,<a href="#B14-cancers-16-02172" class="html-bibr">14</a>,<a href="#B15-cancers-16-02172" class="html-bibr">15</a>,<a href="#B16-cancers-16-02172" class="html-bibr">16</a>,<a href="#B17-cancers-16-02172" class="html-bibr">17</a>,<a href="#B19-cancers-16-02172" class="html-bibr">19</a>,<a href="#B20-cancers-16-02172" class="html-bibr">20</a>,<a href="#B22-cancers-16-02172" class="html-bibr">22</a>,<a href="#B23-cancers-16-02172" class="html-bibr">23</a>,<a href="#B24-cancers-16-02172" class="html-bibr">24</a>,<a href="#B25-cancers-16-02172" class="html-bibr">25</a>,<a href="#B26-cancers-16-02172" class="html-bibr">26</a>,<a href="#B27-cancers-16-02172" class="html-bibr">27</a>,<a href="#B29-cancers-16-02172" class="html-bibr">29</a>].</p>
Full article ">
14 pages, 10663 KiB  
Technical Note
Using the Improved YOLOv5-Seg Network and Sentinel-2 Imagery to Map Glacial Lakes in High Mountain Asia
by Lichen Yin, Xin Wang, Wentao Du, Chengde Yang, Junfeng Wei, Qiong Wang, Dongyu Lei and Jingtao Xiao
Remote Sens. 2024, 16(12), 2057; https://doi.org/10.3390/rs16122057 - 7 Jun 2024
Viewed by 548
Abstract
Continuously monitoring and mapping glacial lake variation is of great importance for determining changes in water resources and potential hazards in alpine cryospheric regions. The semi-automated glacial lake mapping methods used currently are hampered by inherent subjectivity and inefficiency. This study used improved [...] Read more.
Continuously monitoring and mapping glacial lake variation is of great importance for determining changes in water resources and potential hazards in alpine cryospheric regions. The semi-automated glacial lake mapping methods used currently are hampered by inherent subjectivity and inefficiency. This study used improved YOLOv5 strategies to extract glacial lake boundaries from Sentinel-2 imagery. These strategies include using the space-to-depth technique to identify small glacial lakes, and adopting the coordinate attention and the convolution block attention modules to improve mapping performance and adaptability. In terms of glacial lake extraction, the improved YOLOv5-seg network achieved values of 0.95, 0.93, 0.96, and 0.94 for precision (P), recall (R), mAP_0.5, and the F1 score, respectively, indicating an overall improvement in performance of 12% compared to that of the newest YOLOv8 networks. In High Mountain Asia (HMA), 23,108 glacial lakes with a total area of 1847.5 km² were identified in imagery from 2022 using the proposed method. Compared with the use of manual interpretation for lake boundary extraction in test sites of HMA, the proposed method achieved values of 0.89, 0.87, and 0.86 for P, R, and the F1 score, respectively. Our proposed deep learning method has improved accuracy in glacial lake extraction because it can address the challenge represented by frozen or high-turbidity glacial lakes in HMA. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of glacial lakes of different sizes in the glaciered region of HMA (pink). Black squares identify where monthly variations in glacial lakes area were detected.</p>
Full article ">Figure 2
<p>Temporal phase of remote sensing images in High Mountain Asia (143 Sentinel-2 images from 2018 were used to train the deep learning model; 1644 Sentinel-2 images from 2022 were used for glacial lake boundary extraction).</p>
Full article ">Figure 3
<p>Flowchart of glacial lake boundary extraction method.</p>
Full article ">Figure 4
<p>Coordinate attention module (CONV represents a convolution layer, and the kernel size is 1. X AVGPOOL and Y AVGPOOL means the average adaptive pooling of X and Y, respectively. Split represents the splitting of the tensor into 1 <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> H <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> C, 1 <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> W <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> C. W represents the width. H represents the height. C represents the number of channels).</p>
Full article ">Figure 5
<p>Convolution block attention module.</p>
Full article ">Figure 6
<p>Structure of the SPD-Conv module.</p>
Full article ">Figure 7
<p>Network structure adopted in this study (convolution layer 3 <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> 3: convolution kernel is a convolution layer with 3 steps of 2, convolution layer 1 <math display="inline"><semantics> <mrow> <mo>×</mo> </mrow> </semantics></math> 1: convolution kernel is 1, step size is 1).</p>
Full article ">Figure 8
<p>Glacial lake distribution results for the HMA and its subregions in 2022.</p>
Full article ">Figure 9
<p>Contribution of different factors in glacial lake detection (a—RGB, b—synthesis of bands 8, 4, 3, c—synthesis of bands 11, 4, 3, d—adding terrain factors).</p>
Full article ">Figure 10
<p>Relative anomaly of glacial lake area during June–October in HMA in 2022 (orange line represents the average of the relative anomaly of monthly glacial lake area; the left and right limits of the box represent the upper and lower quartiles of the monthly relative anomaly, respectively; and the whisker lines indicate the maximum relative anomaly of the different subregions of HMA).</p>
Full article ">
19 pages, 2537 KiB  
Article
Use of Real-World FHIR Data Combined with Context-Sensitive Decision Modeling to Guide Sentinel Biopsy in Melanoma
by Catharina Lena Beckmann, Georg Lodde, Jessica Swoboda, Elisabeth Livingstone and Britta Böckmann
J. Clin. Med. 2024, 13(11), 3353; https://doi.org/10.3390/jcm13113353 - 6 Jun 2024
Viewed by 591
Abstract
Background: To support clinical decision-making at the point of care, the “best next step” based on Standard Operating Procedures (SOPs) and actual accurate patient data must be provided. To do this, textual SOPs have to be transformed into operable clinical algorithms and [...] Read more.
Background: To support clinical decision-making at the point of care, the “best next step” based on Standard Operating Procedures (SOPs) and actual accurate patient data must be provided. To do this, textual SOPs have to be transformed into operable clinical algorithms and linked to the data of the patient being treated. For this linkage, we need to know exactly which data are needed by clinicians at a certain decision point and whether these data are available. These data might be identical to the data used within the SOP or might integrate a broader view. To address these concerns, we examined if the data used by the SOP is also complete from the point of view of physicians for contextual decision-making. Methods: We selected a cohort of 67 patients with stage III melanoma who had undergone adjuvant treatment and mainly had an indication for a sentinel biopsy. First, we performed a step-by-step simulation of the patient treatment along our clinical algorithm, which is based on a hospital-specific SOP, to validate the algorithm with the given Fast Healthcare Interoperability Resources (FHIR)-based data of our cohort. Second, we presented three different decision situations within our algorithm to 10 dermatooncologists, focusing on the concrete patient data used at this decision point. The results were conducted, analyzed, and compared with those of the pure algorithmic simulation. Results: The treatment paths of patients with melanoma could be retrospectively simulated along the clinical algorithm using data from the patients’ electronic health records. The subsequent evaluation by dermatooncologists showed that the data used at the three decision points had a completeness between 84.6% and 100.0% compared with the data used by the SOP. At one decision point, data on “patient age (at primary diagnosis)” and “date of first diagnosis” were missing. Conclusions: The data needed for our decision points are available in the FHIR-based dataset. Furthermore, the data used at decision points by the SOP and hence the clinical algorithm are nearly complete compared with the data required by physicians in clinical practice. This is an important precondition for further research focusing on presenting decision points within a treatment process integrated with the patient data needed. Full article
(This article belongs to the Section Dermatology)
Show Figures

Figure 1

Figure 1
<p>The entire research process, starting with the initial situation, simulation of the real-world EHR data along the clinical algorithm, and evaluation of data completeness at a specific decision point (green). EHR = electronic health records, SOP = standard operating procedure.</p>
Full article ">Figure 2
<p>Patient flow chart. ICI = immune checkpoint inhibition; TT = targeted therapy; MUP = melanoma of unknown primary; SLNE = sentinel lymph node excision; SOP = standard operating procedure.</p>
Full article ">Figure 3
<p>Graphical representation of the pre-processing procedure used for simulation.</p>
Full article ">Figure 4
<p>Component diagram of the technical set-up for simulation.</p>
Full article ">Figure 5
<p>A snippet (translated into English) of the clinical algorithm [<a href="#B20-jcm-13-03353" class="html-bibr">20</a>,<a href="#B21-jcm-13-03353" class="html-bibr">21</a>] that was presented to physicians to evaluate data needs at the decision point (highlighted in green), here exemplified by decision point DP8 (see <a href="#jcm-13-03353-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 6
<p>Abstract flowchart of the treatment of melanoma stage III-patients according to the SOP document (from DP1 onwards, a distinction is made among patients with an indication for a sentinel biopsy, MUP patients, and patients who only receive guideline-based follow-up care). The sequence of decision points (green, listed in <a href="#jcm-13-03353-t002" class="html-table">Table 2</a>) and the proportionate number of the 67 adjuvantly treated patients from our initial cohort who passed through the individual treatment sections during retrospective simulation (blue) are shown. * 12 patients left the SLNE section because of highly individualized treatment in terms of modeling the clinical algorithm, and one patient left because there was no risk constellation to receive an SLNE. DP = decision point, MUP = melanoma of unknown primary, SLNE = sentinel lymph node excision, SOP = standard operating procedure.</p>
Full article ">
20 pages, 7186 KiB  
Article
Mapping Forest Stock Volume Using Phenological Features Derived from Time-Serial Sentinel-2 Imagery in Planted Larch
by Qianyang Li, Hui Lin, Jiangping Long, Zhaohua Liu, Zilin Ye, Huanna Zheng and Peisong Yang
Forests 2024, 15(6), 995; https://doi.org/10.3390/f15060995 - 6 Jun 2024
Viewed by 594
Abstract
As one of the important types of forest resources, mapping forest stock volume (FSV) in larch (Larix decidua) forests holds significant importance for forest resource management, carbon cycle research, and climate change monitoring. However, the accuracy of FSV mapping using common [...] Read more.
As one of the important types of forest resources, mapping forest stock volume (FSV) in larch (Larix decidua) forests holds significant importance for forest resource management, carbon cycle research, and climate change monitoring. However, the accuracy of FSV mapping using common spectral and texture features is often limited due to their failure in fully capturing seasonal changes and growth cycle characteristics of vegetation. Phenological features can effectively provide essential information regarding the growth status of forests. In this study, multi-temporal Sentinel-2 satellite imagery were initially acquired in the Wangyedian Forest Farm in Chifeng City, Inner Mongolia. Subsequently, various phenological features were extracted from time series variables constructed by Gaussian Process Regression (GPR) using Savitzky–Golay filters, stepwise differentiation, and Fourier transform techniques. The alternative features were further refined through Pearson’s correlation coefficient analysis and the forward selection algorithm, resulting in six groups of optimal subsets. Finally, four models including the Random Forest (RF), K-Nearest Neighbors (KNN), Support Vector Machine (SVM), and Multiple Linear Regression (MLR) algorithms were developed to estimate FSV. The results demonstrated that incorporating phenological features significantly enhanced model performance, with the SVM model exhibiting the best performance—achieving an R2 value of 0.77 along with an RMSE value of 46.36 m3/hm2 and rRMSE value of 22.78%. Compared to models without phenological features, inclusion of these features led to a 0.25 increase in R2 value while reducing RMSE by 10.40 m3/hm2 and rRMSE by 5%. Overall, integration of phenological feature variables not only improves the accuracy of larch forest FSV mapping but also has potential implications for delaying saturation phenomena. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The location of the study area and maps of ground samples: (<b>a</b>) is a map of the location of Inner Mongolia in China; (<b>b</b>) is a map of the location of the forest farm in Chifeng City; (<b>c</b>) is a sample point on the ground in the study area; (<b>d</b>,<b>e</b>) is a close-up photograph of a larch plantation forest.</p>
Full article ">Figure 2
<p>The relationships among DBH, Height and FSV: (<b>a</b>) is the scatterplot of FSV and DBH; (<b>b</b>) is the scatterplot of FSV and Height.</p>
Full article ">Figure 3
<p>Sentinel-2 images for four seasons.</p>
Full article ">Figure 4
<p>Gaussian Process Regression to reconstruct NDVI: (<b>a</b>) Unreconstructed NDVI values; (<b>b</b>) Reconstructed NDVI values.</p>
Full article ">Figure 5
<p>Schematic diagram of time series data reconstructed by S–G filter.</p>
Full article ">Figure 6
<p>S–G filtering to reconstruct NDVI and extract phenology features.</p>
Full article ">Figure 7
<p>Extracting phenological features using time series seasonal differences.</p>
Full article ">Figure 8
<p>Fourier extraction of growing season amplitude variables.</p>
Full article ">Figure 9
<p>The technical roadmap for mapping forest stock volume based on phenological feature data extraction.</p>
Full article ">Figure 10
<p>Distribution of Pearson’s correlation coefficients between different types of features and FSV: (<b>a</b>) Spectral features; (<b>b</b>) Textural features; (<b>c</b>–<b>f</b>) Phenological features by various methods.</p>
Full article ">Figure 11
<p>The scatterplots between measured and predicted FSV using Sentinel-2A images acquired in September.</p>
Full article ">Figure 12
<p>The scatterplots between the measured and predicted FSV using various phenological features.</p>
Full article ">Figure 13
<p>The map of FSV in the planted larch forest of study area.</p>
Full article ">Figure 14
<p>The histograms of accuracy indices derived from various subsets (<b>a</b>) the histograms of R2 derived from various subsets; (<b>b</b>) the histograms of rRMSE derived from various subsets.</p>
Full article ">
27 pages, 4080 KiB  
Review
Satellite Remote Sensing Tools for Drought Assessment in Vineyards and Olive Orchards: A Systematic Review
by Nazaret Crespo, Luís Pádua, João A. Santos and Helder Fraga
Remote Sens. 2024, 16(11), 2040; https://doi.org/10.3390/rs16112040 - 6 Jun 2024
Viewed by 1294
Abstract
Vineyards and olive groves are two of the most important Mediterranean crops, not only for their economic value but also for their cultural and environmental significance, playing a crucial role in global agriculture. This systematic review, based on an adaptation of the 2020 [...] Read more.
Vineyards and olive groves are two of the most important Mediterranean crops, not only for their economic value but also for their cultural and environmental significance, playing a crucial role in global agriculture. This systematic review, based on an adaptation of the 2020 PRISMA statement, focuses on the use of satellite remote sensing tools for the detection of drought in vineyards and olive groves. This methodology follows several key steps, such as defining the approach, selecting keywords and databases, and applying exclusion criteria. The bibliometric analysis revealed that the most frequently used terms included “Google Earth Engine” “remote sensing” “leaf area index” “Sentinel-2”, and “evapotranspiration”. The research included a total of 81 articles published. The temporal distribution shows an increase in scientific production starting in 2018, with a peak in 2021. Geographically, the United States, Italy, Spain, France, Tunisia, Chile, and Portugal lead research in this field. The studies were classified into four categories: aridity and drought monitoring (ADM), agricultural water management (AWM), land use management (LUM), and water stress (WST). Research trends were analysed in each category, highlighting the use of satellite platforms and sensors. Several case studies illustrate applications in vineyards and olive groves, especially in semi-arid regions, focusing on the estimation of evapotranspiration, crop coefficients, and water use efficiency. This article provides a comprehensive overview of the current state of research on the use of satellite remote sensing for drought assessment in grapevines and olive trees, identifying trends, methodological approaches, and opportunities for future research in this field. Full article
(This article belongs to the Special Issue Remote Sensing in Viticulture II)
Show Figures

Figure 1

Figure 1
<p>Average grape production (t) by country between 2018 and 2022 (<b>a</b>); worldwide grape production evolution between 1961 and 2022 (<b>b</b>); and percentage of vineyard plantation area by continent in 2022 (<b>c</b>). Adapted from FAOSTAT.</p>
Full article ">Figure 2
<p>Average olive production (t) by country between 2018 and 2022 (<b>a</b>); worldwide olive production evolution between 1961 and 2022 (<b>b</b>); and percentage of olive plantation area by continent in 2022 (<b>c</b>). Adapted from FAOSTAT.</p>
Full article ">Figure 3
<p>PRISMA flow diagram of the systematic literature review search adapted from Moher et al. [<a href="#B34-remotesensing-16-02040" class="html-bibr">34</a>].</p>
Full article ">Figure 4
<p>Network connection graph between the top 20 most frequently used keywords in the selected studies. The size of each circle corresponds to the frequency of a keyword’s usage, with larger circles indicating higher usage. The top 5 most used keywords are distinguished in red. The proximity between circles, connected by lines, identifies the degree of connection between the corresponding keywords.</p>
Full article ">Figure 5
<p>Temporal distribution of articles included in the systematic review by publication year (<b>a</b>) and the spatial distribution by country (<b>b</b>).</p>
Full article ">Figure 6
<p>Distribution of studies expressed as a percentage, categorised by crop type.</p>
Full article ">Figure 7
<p>Percentage of publications between 1 January 2003 and 30 December 2023, according to categorical classification by country (<b>a</b>) and by year (<b>b</b>). ADM: aridity and drought monitoring; AWM: agricultural water management; LUM: land use management; and WST: water stress. AF: Afghanistan; AU: Australia; CL: Chile; CN: China; FR: France; GR: Greece; IT: Italy; LB: Lebanon; MD: Moldova; MA: Morocco; PT: Portugal; KSA: Saudi Arabia (Kingdom of); ES: Spain; TN: Tunisia; TR: Türkiye; USA: United States of America.</p>
Full article ">Figure 8
<p>Percentage distribution of studies by satellite platform (<b>a</b>) and by sensor (<b>b</b>) based on their respective categorical classification. ADM: aridity and drought monitoring; AWM: agricultural water management; LUM: land use management; and WST: water stress. Note that one study can appear several times, as it may include more than one index.</p>
Full article ">Figure 9
<p>The distribution of studies by the use of indices according to their respective categorical classifications. ADM: aridity and drought monitoring; AWM: agricultural water management; LUM: land use management; and WST: water stress. Others correspond to the sum of vegetation indices that have been used only once. Note that one study can appear several times, as it may include more than one index.</p>
Full article ">
30 pages, 18476 KiB  
Article
Mapping Maize Evapotranspiration with Two-Source Land Surface Energy Balance Approaches and Multiscale Remote Sensing Imagery Pixel Sizes: Accuracy Determination toward a Sustainable Irrigated Agriculture
by Edson Costa-Filho, José L. Chávez and Huihui Zhang
Sustainability 2024, 16(11), 4850; https://doi.org/10.3390/su16114850 - 6 Jun 2024
Viewed by 753
Abstract
This study evaluated the performance of remote sensing (RS) algorithms for the estimation of actual maize evapotranspiration (ETa) using different spaceborne, airborne, and proximal multispectral data in a semi-arid climate region to identify the optimal platform that provides the best ET [...] Read more.
This study evaluated the performance of remote sensing (RS) algorithms for the estimation of actual maize evapotranspiration (ETa) using different spaceborne, airborne, and proximal multispectral data in a semi-arid climate region to identify the optimal platform that provides the best ETa estimates to improve irrigation water management and help make irrigated agriculture sustainable. The RS platforms used in the study included Landsat-8 (30 m pixel spatial resolution), Sentinel-2 (10 m), Planet CubeSat (3 m), multispectral radiometer or MSR (1 m), and a small uncrewed aerial system or sUAS (0.03 m). Two-source surface energy balance (TSEB) models, implementing the series and parallel surface resistance approaches, were used in this study to estimate hourly maize ETa. The data used in this study were obtained from two maize research sites in Greeley and Fort Collins, CO, USA, in 2020 and 2021. Each research site had different irrigation systems. The Greeley site had a subsurface drip system, while the Fort Collins site had surface irrigation (furrow). Maize ETa predictions were compared to observed maize ETa data from an eddy covariance system installed at each research site. Results indicated that the MSR5 proximal platform (1 m) provided optimal RS data for the TSEB algorithms. The MSR5 “point-based” nadir-looking surface reflectance data and surface radiometric temperature combination resulted in the smallest error when predicting hourly (mm/h) maize ETa. The mean bias and root mean square errors (MBE and RMSE, respectively), when predicting maize hourly ETa using the MSR5 sensor data, were equal to −0.02 (−3%) ± 0.07 (11%) mm/h MBE ± RMSE and −0.02 (−3%) ± 0.09 (14%) mm/h for the TSEB parallel and series approaches, respectively. The poorest performance, when predicting hourly TSEB maize ETa, was from Landsat-8 (30 m) multispectral data combined with its original thermal data, since the errors were −0.03 (−5%) ± 0.16 (29%) mm/h and −0.07 (−13%) ± 0.15 (29%) mm/h for the TSEB parallel and series approaches, respectively. These results indicate the need to develop methods to improve the quality of the RS data from sub-optimal platforms/sensors/scales/calibration to further advance sustainable irrigation water management. Full article
Show Figures

Figure 1

Figure 1
<p>False-color image of the LIRF research site near Greeley, CO, USA. The study maize fields were Fields W and E located in the southeast corner of the research farm.</p>
Full article ">Figure 2
<p>Plant variety map at LIRF (Fields W and E) in 2021. Most of the area was occupied by the P0157AMXT maize variety.</p>
Full article ">Figure 3
<p>The 2020–2021 LIRF experiment design.</p>
Full article ">Figure 4
<p>RGB (red–green–blue) map of the IIC research fields (<b>a</b>) and the maize varieties planted in 2021 (<b>b</b>). The study maize fields were Fields F and D. Areas in green are vegetation surfaces.</p>
Full article ">Figure 5
<p>EC systems were installed at the LIRF (<b>a</b>) and IIC (<b>b</b>) sites in 2020 and 2021 at 3.5 m AGS. (<b>a</b>) courtesy of Jon Altenhofen.</p>
Full article ">Figure 6
<p>Two-dimensional EC footprint (yellow areas) at LIRF maize fields in 2020 (<b>a</b>) and 2021 (<b>b</b>).</p>
Full article ">Figure 7
<p>Two-dimensional EC footprint (yellow areas) at the IIC maize field F in 2020 (<b>a</b>) and 2021 (<b>b</b>).</p>
Full article ">Figure 8
<p>Scatter plots (1:1 line) and error analysis results regarding the TSEB parallel (TSEB<sub>par</sub>) maize hourly ET<sub>a</sub> modeling results for the combined LIRF and IIC 2020–2021 data. The sample size (<span class="html-italic">n</span>) of each platform is indicated in the figure.</p>
Full article ">Figure 9
<p>Scatter plots (1:1 line) and error analysis results regarding the TSEB series (TSEB<sub>ser</sub>) maize hourly ET<sub>a</sub> modeling results for the combined LIRF and IIC 2020–2021 data. The sample size (<span class="html-italic">n</span>) of each platform is indicated in the figure.</p>
Full article ">Figure 10
<p>Error analysis of the SEB fluxes using the TSEB series (TSEB<sub>ser</sub>) algorithm and LIRF and IIC 2020–2021 data combined.</p>
Full article ">Figure 11
<p>Error analysis of the SEB fluxes using the TSEB parallel (TSEB<sub>par</sub>) algorithm and LIRF and IIC 2020–2021 data combined.</p>
Full article ">
10 pages, 3658 KiB  
Article
Sentinel Lymph Node Mapping by Retroperitoneal vNOTES for Uterus-Confined Malignancies: A Standardized 10-Step Approach
by Daniela Huber and Yannick Hurni
Cancers 2024, 16(11), 2142; https://doi.org/10.3390/cancers16112142 - 5 Jun 2024
Viewed by 541
Abstract
(1) Background: Sentinel lymph node (SLN) mapping represents an accurate and feasible technique for the surgical staging of endometrial and cervical cancer. This is commonly performed by conventional laparoscopy or robotic-assisted laparoscopy, but in recent years, a new retroperitoneal transvaginal natural orifice transluminal [...] Read more.
(1) Background: Sentinel lymph node (SLN) mapping represents an accurate and feasible technique for the surgical staging of endometrial and cervical cancer. This is commonly performed by conventional laparoscopy or robotic-assisted laparoscopy, but in recent years, a new retroperitoneal transvaginal natural orifice transluminal endoscopic surgery (vNOTES) approach has been described and developed by Jan Baekelandt. This technique provides easy visualization of lymphatic afferent vessels and pelvic lymph nodes, early SLN assessment, and a coherent mapping methodology following the lymphatic flow from caudal to cranial. However, only a few publications have reported it. Following the IDEAL (Idea Development Exploration Assessment Long-term follow-up) framework, research concerning this technique is in Stage 2a, with only small case series as evidence of its feasibility. Its standardized description appears necessary to provide the surgical homogeneity required to move further. (2) Methods: Description of a standardized approach for retroperitoneal pelvic SLN mapping by vNOTES. (3) Results: We describe a 10-step approach to successfully perform retroperitoneal vNOTES SLN mapping, including pre-, intra-, and postoperative management. (4) Conclusions: This IDEAL Stage 2a study could help other surgeons approach this new technique, and it proposes a common methodology necessary for evolving through future IDEAL Stage 2b (multi-center studies) and Stage 3 (randomized controlled trials) studies. Full article
Show Figures

Figure 1

Figure 1
<p>Retroperitoneal vNOTES view demonstrating the left pelvic anatomical structures.</p>
Full article ">Figure 2
<p>Retroperitoneal vNOTES view demonstrating the right pelvic anatomical structures.</p>
Full article ">Figure 3
<p>Retroperitoneal vNOTES view demonstrating SLN mapping to the right external iliac region.</p>
Full article ">Figure 4
<p>Demonstration of an SLN harvested by retroperitoneal vNOTES from the right external iliac region.</p>
Full article ">
31 pages, 10514 KiB  
Article
Google Earth Engine and Machine Learning for Flash Flood Exposure Mapping—Case Study: Tetouan, Morocco
by EL Mehdi SELLAMI and Hassan Rhinane
Geosciences 2024, 14(6), 152; https://doi.org/10.3390/geosciences14060152 - 4 Jun 2024
Cited by 1 | Viewed by 1592
Abstract
Recently, the earth’s climate has changed considerably, leading to several hazards, including flash floods (FFs). This study aims to introduce an innovative approach to mapping and identifying FF exposure in the city of Tetouan, Morocco. To address this problem, the study uses different [...] Read more.
Recently, the earth’s climate has changed considerably, leading to several hazards, including flash floods (FFs). This study aims to introduce an innovative approach to mapping and identifying FF exposure in the city of Tetouan, Morocco. To address this problem, the study uses different machine learning methods applied to remote sensing imagery within the Google Earth Engine (GEE) platform. To achieve this, the first phase of this study was to map land use and land cover (LULC) using Random Forest (RF), a Support Vector Machine (SVM), and Classification and Regression Trees (CART). By comparing the results of five composite methods (mode, maximum, minimum, mean, and median) based on Sentinel images, LULC was generated for each method. In the second phase, the precise LULC was used as a related factor to others (Stream Power Index (SPI), Topographic Position Index (TPI), Slope, Profile Curvature, Plan Curvature, Aspect, Elevation, and Topographic Wetness Index (TWI)). In addition to 2024 non-flood and flood points to predict and detect FF susceptibility, 70% of the dataset was used to train the model by comparing different algorithms (RF, SVM, Logistic Regression (LR), Multilayer Perceptron (MLP), and Naive Bayes (NB)); the rest of the dataset (30%) was used for evaluation. Model performance was evaluated by five-fold cross-validation to assess the model’s ability on new data using metrics such as precision, score, kappa index, recall, and the receiver operating characteristic (ROC) curve. In the third phase, the high FF susceptibility areas were analyzed for two-way validation with inundated areas generated from Sentinel-1 SAR imagery with coherent change detection (CDD). Finally, the validated inundation map was intersected with the LULC areas and population density for FF exposure and assessment. The initial results of this study in terms of LULC mapping showed that the most appropriate method in this research region is the use of an SVM trained on a mean composite. Similarly, the results of the FF susceptibility assessment showed that the RF algorithm performed best with an accuracy of 96%. In the final analysis, the FF exposure map showed that 2465 hectares were affected and 198,913 inhabitants were at risk. In conclusion, the proposed approach not only allows us to assess the impact of FF in this study area but also provides a versatile approach that can be applied in different regions around the world and can help decision-makers plan FF mitigation strategies. Full article
(This article belongs to the Special Issue Flood Risk Reduction)
Show Figures

Figure 1

Figure 1
<p>Study area: (<b>1</b>) North Africa, Morocco; (<b>2</b>) Tanger<tt>–</tt>Tetouan<tt>–</tt>Al Hoceima region; (<b>3</b>) Tetouan Province; (<b>4</b>) Tetouan city.</p>
Full article ">Figure 2
<p>The database used: (<b>A</b>) an overview of the study area’s satellite imagery, (<b>B</b>) the elevation distribution based on the digital elevation model (DEM) of the study area, (<b>C</b>) the population density repartition in the study area, (<b>D</b>) the flooded (red points) and non-flooded (green points) locations in the study area.</p>
Full article ">Figure 3
<p>The proposed methodology flowchart: (<b>1</b>) Mapping LULC; (<b>2</b>) FF susceptibility mapping and forecasting; (<b>3</b>) FF extent mapping; (<b>4</b>) FF exposure.</p>
Full article ">Figure 4
<p>Flash flood conditioning factors in the study area: (<b>A</b>) Geographical distribution of flooded and non-flooded locations from the field survey, (<b>B</b>) LU/LC, (<b>C</b>) SPI, (<b>D</b>) Plan curvature.</p>
Full article ">Figure 5
<p>Flash flood conditioning factors in the study area: (<b>A</b>) Profile curvature, (<b>B</b>) TPI, (<b>C</b>) TWI, (<b>D</b>) Elevation, (<b>E</b>) Slope, and (<b>F</b>) Aspect.</p>
Full article ">Figure 6
<p>Synthetic Aperture Radar (SAR) images showing the study area before (<b>left</b>) and during (<b>right</b>) the flash flood.</p>
Full article ">Figure 7
<p>Land use and land cover classification of the study area (<b>1</b>–<b>15</b>) for each composite image ((<b>A</b>) Max; (<b>B</b>) Mean; (<b>C</b>) Median; (<b>D</b>) Min; (<b>E</b>) Mode), generated using machine learning algorithms ((<b>I</b>) SVM, (<b>II</b>) RF, (<b>III</b>) CART).</p>
Full article ">Figure 8
<p>The influence of the choice of dataset on classification performance: Evaluation of the impact of different datasets (max, mean, median, min, and mode) on LU/LC classification performance using overall accuracy and the Kappa index for each machine learning algorithm used (SVM, RF, and CART).</p>
Full article ">Figure 9
<p>Flash Flood susceptibility maps generated using machine learning algorithms: Support Vector Machine (SVM), Logistic Regression (LR), Random Forest (RF), Naive Bayes (NB), and Multilayer Perceptron (MLP). Each map categorizes the study area into five levels of flood susceptibility: highest, high, moderate, low, and lowest.</p>
Full article ">Figure 10
<p>Comparison of the performance of the machine learning algorithms used (SVM, LR, RF, NB, and MLP) for flash flood susceptibility using precision, recall, F1 score, and Kappa.</p>
Full article ">Figure 11
<p>Performance comparison of the machine learning algorithms used (SVM, LR, RF, NB, and MLP) using Receiver operating characteristic (ROC) curves.</p>
Full article ">Figure 12
<p>Pairwise correlation matrix of variables used in the study: LU/LC, aspect, TWI, TPI, slope, profile curvature, plan curvature, X, Y, SPI, flood, and non-flood.</p>
Full article ">Figure 13
<p>Comparison of the relative importance of the different factors used for flash food susceptibility mapping, as identified by Random Forest (RF) (<b>A</b>) and Multilayer Perceptron (MLP) (<b>B</b>).</p>
Full article ">Figure 14
<p>Flash flooding extent of the study area using Synthetic Aperture Radar (SAR) data. Flash flood areas are shown in blue.</p>
Full article ">Figure 15
<p>Superposition of high flash flood susceptibility and SAR flash flood data. Areas of high flash flood susceptibility are shown in yellow. Areas detected as inundated using Synthetic Aperture Radar (SAR) data are shown in blue.</p>
Full article ">Figure 16
<p>Overview of the Google Earth Engine application showing exposure to flash floods in the study area. Affected forest areas are shown in dark green, affected vegetation in light green, affected urban areas in red, and affected barren areas in yellow.</p>
Full article ">Figure 17
<p>Distribution of the population potentially exposed to flash flooding in the study area based on the population density.</p>
Full article ">Figure 18
<p>Flash flood exposure in the study areas: Colors represent affected areas: dark green (forest), light green (vegetation), red (urban), and yellow (barren land).</p>
Full article ">
17 pages, 14166 KiB  
Article
Source Models of the 2016 and 2022 Menyuan Earthquakes and Their Tectonic Implications Revealed by InSAR
by Xixuan Bai, Bingqiang Zhang, Aizhi Guo, Yi Yan, Hao Xu, Xiaoya Bian, Shuwen Zhan and Jiangcheng Chen
Sensors 2024, 24(11), 3622; https://doi.org/10.3390/s24113622 - 4 Jun 2024
Viewed by 317
Abstract
The Haiyuan fault system plays a crucial role in accommodating the eastward expansion of the Tibetan Plateau (TP) and is currently slipping at a rate of several centimeters per year. However, limited seismic activities have been observed using geodetic techniques in this area, [...] Read more.
The Haiyuan fault system plays a crucial role in accommodating the eastward expansion of the Tibetan Plateau (TP) and is currently slipping at a rate of several centimeters per year. However, limited seismic activities have been observed using geodetic techniques in this area, impeding the comprehensive investigation into regional tectonics. In this study, the geometric structure and source models of the 2022 Mw 6.7 and the 2016 Mw 5.9 Menyuan earthquakes were investigated using Sentinel-1A SAR images. By implementing an atmospheric error correction method, the signal-to-noise ratio of the 2016 interferometric synthetic aperture radar (InSAR) coseismic deformation field was significantly improved, enabling InSAR observations with higher accuracy. The results showed that the reliability of the source models for those events was improved following the reduction in observation errors. The Coulomb stress resulting from the 2016 event may have promoted the strike-slip movement of the western segment of the Lenglongling fault zone, potentially expediting the occurrence of the 2022 earthquake. The coseismic slip distribution and the spatial distribution of aftershocks of the 2022 event suggested that the seismogenic fault may connect the western segment of the Lenglongling fault (LLLF) and the eastern segment of the Tuolaishan fault (TLSF). Additionally, the western segment of the surface rupture zone of the northern branch may terminate in the secondary branch close to the Sunan-Qilian fault (SN-QL) strike direction, and the earthquake may have triggered deep aftershocks and accelerated stress release within the deep seismogenic fault. Full article
(This article belongs to the Section Remote Sensors)
Show Figures

Figure 1

Figure 1
<p>Tectonic settings of the Haiyuan fault system. (<b>a</b>) The movement trend of the Tibetan Plateau under the compression of the Indian plate; the blue rectangle is the range of (<b>b</b>), and the red arrows are the GPS velocities of the permanent stations [<a href="#B9-sensors-24-03622" class="html-bibr">9</a>]. (<b>b</b>) The blue and red beach balls represent the focal mechanisms of the 2016 and 2022 events from the United States Geological Survey (USGS), respectively. The black lines represent active faults modified from [<a href="#B9-sensors-24-03622" class="html-bibr">9</a>,<a href="#B10-sensors-24-03622" class="html-bibr">10</a>], and the solid black dots denote historical earthquakes (1920~2022) from the USGS catalog. Blue arrows indicate the Global Positioning System (GPS) velocities of the roving stations, and red arrows are the GPS velocities of the permanent stations [<a href="#B11-sensors-24-03622" class="html-bibr">11</a>]. The green and yellow rectangles indicate the Sentinel-1 ascending and descending track SAR coverages. The red rectangle represents the range of Subfigure (<b>c</b>), where the pentagram represents the epicenter of the two main earthquakes after relocation, and the solid point indicates the magnitude and spatial location [<a href="#B12-sensors-24-03622" class="html-bibr">12</a>,<a href="#B13-sensors-24-03622" class="html-bibr">13</a>].</p>
Full article ">Figure 2
<p>The function models of atmospheric delay correction. The blue scatter represents the distribution of noise and elevation, and the yellow solid line represents the function based on the fit between noise and elevation; (<b>a</b>) is the function model of the ascending track (T128), (<b>b</b>) is the function model of the descending track (T33).</p>
Full article ">Figure 3
<p>Coseismic surface deformations of the Menyuan earthquakes in 2016 and 2022. (<b>a</b>,<b>d</b>) Coseismic interferograms of the 2016 earthquake, (<b>b</b>,<b>e</b>) the line-of-sight (LOS) displacements for radars obtained from these interferograms, (<b>c</b>,<b>f</b>) line-of-sight (LOS) displacements after correcting atmospheric delays. (<b>g</b>–<b>i</b>) The coseismic interferogram of the 2022 earthquake, (<b>j</b>–<b>l</b>) the line-of-sight (LOS) displacements obtained from these interferograms, and (<b>m</b>–<b>o</b>) the deformations after masking the area of low coherence. The black solid line represents the fault trace of the modified reference [<a href="#B17-sensors-24-03622" class="html-bibr">17</a>,<a href="#B18-sensors-24-03622" class="html-bibr">18</a>], the red beach ball represents the focal mechanism of the 2022 event, and the blue ball represents the 2016 event. The rectangle formed by the blue and red lines is the vertical projection of the fault plane on the surface, where the red line represents the top boundary of the fault plane.</p>
Full article ">Figure 4
<p>Coseismic slip distribution corresponding to events in 2016; the square represents a discrete sub-fault, and the black arrow represents the slip vector; (<b>a</b>,<b>b</b>) show the coseismic slip distribution before atmospheric delay correction; (<b>c</b>,<b>d</b>) display the coseismic slip distribution after atmospheric delay correction.</p>
Full article ">Figure 4 Cont.
<p>Coseismic slip distribution corresponding to events in 2016; the square represents a discrete sub-fault, and the black arrow represents the slip vector; (<b>a</b>,<b>b</b>) show the coseismic slip distribution before atmospheric delay correction; (<b>c</b>,<b>d</b>) display the coseismic slip distribution after atmospheric delay correction.</p>
Full article ">Figure 5
<p>Coseismic slip distribution corresponding to events in 2022; the square represents a discrete sub-fault, and the black arrow represents a slip vector; (<b>a</b>) is the coseismic slip distribution model 1, determined by the initial InSAR observations; (<b>b</b>) is the coseismic slip distribution model 2, obtained by the coseismic deformations after masking the low-coherence area; (<b>c</b>) is model 3 and replaces the F2–F5 segments in (<b>b</b>) with the F2 segment.</p>
Full article ">Figure 6
<p>Coseismic deformation and residual based on coseismic slip distribution forward modeling of the 2022 Menyuan earthquake; (<b>a</b>–<b>c</b>) illustrate the InSAR observations, (<b>d</b>–<b>f</b>) show the simulation results, and (<b>g</b>–<b>i</b>) display the residuals.</p>
Full article ">Figure 7
<p>Coulomb stress changes caused by the two earthquakes; (<b>a</b>–<b>c</b>) Coulomb stress changes at 5 km, 10 km, and 15 km caused by the 2016 earthquake; (<b>d</b>–<b>f</b>) Coulomb stress changes caused by the coseismic slip distribution in <a href="#sensors-24-03622-f005" class="html-fig">Figure 5</a>a; (<b>g</b>–<b>i</b>) those corresponding to coseismic slip distribution in <a href="#sensors-24-03622-f005" class="html-fig">Figure 5</a>b; (<b>j</b>–<b>l</b>) those corresponding to <a href="#sensors-24-03622-f005" class="html-fig">Figure 5</a>c. The depth from left to right is 5 km, 10 km, and 15 km. The solid black lines represent the fault traces modified based on the literature [<a href="#B9-sensors-24-03622" class="html-bibr">9</a>,<a href="#B10-sensors-24-03622" class="html-bibr">10</a>] (Deng et al., 2003; Zhang et al., 2002).</p>
Full article ">
18 pages, 2812 KiB  
Article
Analysing Spatiotemporal Variability of Chlorophyll-a Concentration and Water Surface Temperature in Coastal Lagoons of the Ebro Delta (NW Mediterranean Sea, Spain)
by Lara Talavera, José Antonio Domínguez-Gómez, Nuria Navarro and Inmaculada Rodríguez-Santalla
J. Mar. Sci. Eng. 2024, 12(6), 941; https://doi.org/10.3390/jmse12060941 - 3 Jun 2024
Viewed by 446
Abstract
Coastal lagoons are highly productive transitional water bodies threatened by human factors and vulnerable to global climate change effects. Monitoring biophysical parameters in these ecosystems is crucial for their preservation. In this work, we used Sentinel-2 and Landsat imagery combined with in situ [...] Read more.
Coastal lagoons are highly productive transitional water bodies threatened by human factors and vulnerable to global climate change effects. Monitoring biophysical parameters in these ecosystems is crucial for their preservation. In this work, we used Sentinel-2 and Landsat imagery combined with in situ data to (1) develop preliminary algorithms for retrieving the Chl-a concentration and water surface temperature of six lagoons located in the Ebro Delta (NE Mediterranean Sea, Spain), and to (2) compute maps and trend lines for analysing their spatiotemporal evolution from 2015 to 2022. Our findings showed that the algorithms’ accuracy ranged from 72% to 78% and had limited potential under high Chl-a concentration regimes. Even so, they revealed the lagoons’ trophic status, usual fluctuations, and deviations of both parameters attributed to seasonal (i.e., light and temperature) and short-term physical (i.e., winds) forcing, as well as valuable spatial patterns potentially useful for conservation efforts and land use planning. Future work will focus on the acquisition of a larger in situ data sample under a range of environmental conditions to improve the algorithms’ robustness, which in turn will allow the investigation of natural and human factors controlling the dynamics of the two investigated parameters. Full article
(This article belongs to the Section Marine Environmental Science)
Show Figures

Figure 1

Figure 1
<p>Location of the Ebro Delta habitats, land use types (CORINE 2018), and the two bays (1 and 6) and four coastal lagoons (2, 3, 4 and 5) of interest (Background image: Sentinel S2A-L1C image in November 2022). The in situ measurement points from the ACA (Catalan Water Agency) and ZOCOMAR research group are highlighted with red and orange triangles, respectively.</p>
Full article ">Figure 2
<p>Performance of the algorithms to extract: (<b>a</b>) Chl<span class="html-italic">-a</span> concentration from El Fangar and Els Alfacs bays (Algorithm A1), (<b>b</b>) Chl<span class="html-italic">-a</span> concentration from Bassa de L’Estella, El Garxal, Calaixos de Buda, and La Tancada coastal lagoons (Algorithm A2), and (<b>c</b>) water surface temperature in all the lagoons (Note: the displayed lines correspond to the 1:1 lines, which help to visually understand the differences between the measured and modelled data. The regression lines were also calculated and displayed R<sup>2</sup> values of 0.95, 0.96, and 0.90, for the A1, A2, and temperature algorithms, respectively).</p>
Full article ">Figure 3
<p>Mean Chl<span class="html-italic">-a</span> concentration and standard deviation for the two bays (<b>a</b>) and the four coastal lagoons (<b>b</b>) during the period of study (Note: W = Winter, Sp = Spring, Su = Summer, and A = Autumn).</p>
Full article ">Figure 4
<p>Maps displaying Chl<span class="html-italic">-a</span> concentration patterns for the bays and coastal lagoons analysed in 2021.</p>
Full article ">Figure 5
<p>Modelled water surface temperature trend lines from 2013 to 2022 in (<b>a</b>) El Fangar, (<b>b</b>) Bassa de L’Estella, (<b>c</b>) El Garxal, (<b>d</b>) Calaixos de Buda, (<b>e</b>) La Tancada, and (<b>f</b>) Els Alfacs.</p>
Full article ">Figure 6
<p>Maps displaying water surface temperature patterns for the bays and coastal lagoons analysed in 2018.</p>
Full article ">Figure 7
<p>Maps displaying water surface temperature patterns for the bays and coastal lagoons analysed in 2022.</p>
Full article ">
Back to TopTop