Nothing Special   »   [go: up one dir, main page]

Next Issue
Volume 15, April-2
Previous Issue
Volume 15, March-2
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sens., Volume 15, Issue 7 (April-1 2023) – 237 articles

Cover Story (view full-size image): The Landsat 8 and 9 Underfly Event occurred in November 2021, where L9 flew beneath L8 in the final stages before settling in its final orbiting path. An initial analysis, dubbed “Phase 1”, was performed on the images taken during this event, which resulted in a cross-calibration with uncertainties estimated to be less than 1%. This level of precision was due in part to the near-identical sensors aboard each instrument as well as the underfly event itself, which allowed the sensors to take nearly the exact same image at nearly the exact same time. There were three forms of uncertainty established in Phase 1: geometric, spectral, and angular. This paper covers Phase 2 of the underfly analysis, which made several modifications to the Phase 1 process. The results here were used by USGS EROS to update the L9 calibration at the end of 2022. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
26 pages, 12137 KiB  
Article
GF-1/6 Satellite Pixel-by-Pixel Quality Tagging Algorithm
by Xin Fan, Hao Chang, Lianzhi Huo and Changmiao Hu
Remote Sens. 2023, 15(7), 1955; https://doi.org/10.3390/rs15071955 - 6 Apr 2023
Cited by 1 | Viewed by 1807
Abstract
The Landsat and Sentinel series satellites contain their own quality tagging data products, marking the source image pixel by pixel with several specific semantic categories. These data products generally contain categories such as cloud, cloud shadow, land, water body, and snow. Due to [...] Read more.
The Landsat and Sentinel series satellites contain their own quality tagging data products, marking the source image pixel by pixel with several specific semantic categories. These data products generally contain categories such as cloud, cloud shadow, land, water body, and snow. Due to the lack of mid-wave and thermal infrared bands, the accuracy of traditional cloud detection algorithm is unstable when facing Chinese Gaofen-1/6 (GF-1/6) data. Moreover, it is challenging to distinguish clouds from snow. In order to produce GF-1/6 satellite pixel-by-pixel quality tagging data products, this paper builds a training sample set of more than 100,000 image pairs, primarily using Sentinel-2 satellite data. Then, we adopt the Swin Transformer model with a self-attention mechanism for GF-1/6 satellite image quality tagging. Experiments show that the model’s overall accuracy reaches the level of Fmask v4.6 with more than 10,000 training samples, and the model can distinguish between cloud and snow correctly. Our GF-1/6 quality tagging algorithm can meet the requirements of the “Analysis Ready Data (ARD) Technology Research for Domestic Satellite” project. Full article
(This article belongs to the Special Issue Gaofen 16m Analysis Ready Data)
Show Figures

Figure 1

Figure 1
<p>Comparison of fixed mapping transformation (<b>a</b>,<b>b</b>) and dynamic stretching transformation (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 2
<p>Comparison of Sentinel-2 and GF-1 sample image.</p>
Full article ">Figure 3
<p>Example of sample image and sample label.</p>
Full article ">Figure 4
<p>Flowchart of training sample production.</p>
Full article ">Figure 5
<p>Map of training sample distribution and quantity.</p>
Full article ">Figure 6
<p>The architecture of the Swin Transformer model.</p>
Full article ">Figure 7
<p>Model training process flowchart.</p>
Full article ">Figure 8
<p>GF-1/6 quality tagging algorithm flowchart.</p>
Full article ">Figure 9
<p>Offset chunking process to solve the seam problem.</p>
Full article ">Figure 10
<p>Example of water misdetection and the corresponding correction.</p>
Full article ">Figure 11
<p>Typical visual examples of Sentinel-2 sample images and quality tagging masks (512 × 512) produced by Swin-L. (<b>a</b>–<b>f</b>) Cloud and cloud shadow detection. (<b>g</b>,<b>h</b>) Snow detection.</p>
Full article ">Figure 11 Cont.
<p>Typical visual examples of Sentinel-2 sample images and quality tagging masks (512 × 512) produced by Swin-L. (<b>a</b>–<b>f</b>) Cloud and cloud shadow detection. (<b>g</b>,<b>h</b>) Snow detection.</p>
Full article ">Figure 12
<p>Comparison of Swin-L and Fmask quality tagging result (1). (<b>a</b>) RGB source image. (<b>b</b>) Swin-L label. (<b>c</b>) Fmask label.</p>
Full article ">Figure 13
<p>Comparison of Swin-L and Fmask quality tagging result (2). (<b>a</b>) RGB source image. (<b>b</b>) Swin-L label. (<b>c</b>) Fmask label.</p>
Full article ">Figure 14
<p>Comparison of Swin-L and Fmask quality tagging result (3). (<b>a</b>) RGB source image. (<b>b</b>) Swin-L label. (<b>c</b>) Fmask label.</p>
Full article ">Figure 15
<p>Typical visual examples of GF-1/6 sample images and quality tagging masks (512 × 512) produced by Swin-L. (<b>a</b>–<b>c</b>) Cloud and cloud shadow detection. (<b>d</b>–<b>f</b>) Snow detection.</p>
Full article ">Figure 15 Cont.
<p>Typical visual examples of GF-1/6 sample images and quality tagging masks (512 × 512) produced by Swin-L. (<b>a</b>–<b>c</b>) Cloud and cloud shadow detection. (<b>d</b>–<b>f</b>) Snow detection.</p>
Full article ">Figure 16
<p>Example GF-1 WFV image and quality tagging masks produced by different training sample numbers of Swin-L models. (<b>a</b>) RGB source image; (<b>b</b>) 2k samples based on Swin-L mask; (<b>c</b>) 5k samples based on Swin-L mask; (<b>d</b>)10k samples based on Swin-L mask.</p>
Full article ">Figure 17
<p>Examples of GF-1 WFV full images and quality tagging masks produced by Swin-L. (<b>a</b>) RGB source image. (<b>b</b>) Swin-L label.</p>
Full article ">Figure A1
<p>Example of training samples requiring manual refinement.</p>
Full article ">Figure A2
<p>The Swin-L’s mIoU and mAcc performance with different training sample data volume during training progress.</p>
Full article ">Figure A3
<p>More examples of GF-1/6 WFV full images and quality tagging masks produced by Swin-L. (<b>a</b>) W121.1_N47.6_20170724. (<b>b</b>) W123.5_N45.9_20201123. (<b>c</b>) W121.7_N45.9_20201025. (<b>d</b>) W123.5_N49.2_20201005. (<b>e</b>) W122.2_N47.6_20150418. (<b>f</b>) W122.9_N45.9_20210217. (<b>g</b>) W122.3_N49.3_20160818. (<b>h</b>) W123.5_N47.6_20210612.</p>
Full article ">
15 pages, 5218 KiB  
Technical Note
Terrain Self-Similarity-Based Transformer for Generating Super Resolution DEMs
by Xin Zheng, Zelun Bao and Qian Yin
Remote Sens. 2023, 15(7), 1954; https://doi.org/10.3390/rs15071954 - 6 Apr 2023
Cited by 6 | Viewed by 2227
Abstract
High-resolution digital elevation models (DEMs) are important for relevant geoscience research and practical applications. Compared with traditional hardware-based methods, super-resolution (SR) reconstruction techniques are currently low-cost and feasible methods used for obtaining high-resolution DEMs. Single-image super-resolution (SISR) techniques have become popular in DEM [...] Read more.
High-resolution digital elevation models (DEMs) are important for relevant geoscience research and practical applications. Compared with traditional hardware-based methods, super-resolution (SR) reconstruction techniques are currently low-cost and feasible methods used for obtaining high-resolution DEMs. Single-image super-resolution (SISR) techniques have become popular in DEM SR in recent years. However, DEM super-resolution has not yet utilized reference-based image super-resolution (RefSR) techniques. In this paper, we propose a terrain self-similarity-based transformer (SSTrans) to generate super-resolution DEMs. It is a reference-based image super-resolution method that automatically acquires reference images using terrain self-similarity. To verify the proposed model, we conducted experiments on four distinct types of terrain and compared them to the results from the bicubic, SRGAN, and SRCNN approaches. The experimental results show that the SSTrans method performs well in all four terrains and has outstanding advantages in complex and uneven surface terrains. Full article
(This article belongs to the Section AI Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The self-similarity transformer workflow. DEM-HR refers to the high-resolution DEM data used for comparison with SR data. DEM-LR refers to the low-resolution DEM data obtained after downsampling DEM-HR as input. DEM-Ref refers to the reference data obtained using self-similarity. DEM-LR↑ refers to the data obtained by upsampling DEM-LR, while DEM-Ref↓↑ refers to the data obtained by downsampling and upsampling the reference data. DEM-LR↑ and DEM-Ref↓↑ will be used to calculate the correlation between the low-resolution image and the reference image.</p>
Full article ">Figure 2
<p>Data preprocessing flow. DEM-HR refers to the high-resolution DEM data used for comparison with SR data. DEM-LR refers to the low-resolution DEM data obtained after downsampling DEM-HR as input. DEM-Ref refers to the reference data obtained using self-similarity. DEM-LR↑ refers to the data obtained by upsampling DEM-LR, while DEM-Ref↓↑ refers to the data obtained by downsampling and upsampling the reference data. We will use DEM-LR↑ and DEM-Ref↓↑ to calculate the correlation between the low-resolution image and the reference image.</p>
Full article ">Figure 3
<p>SSTrans structure. F, V, K, Q are the features of DEM-LR, DEM-Ref, DEM-Ref↓↑, and DEM-Ref extracted by the residual network. P, W are the position matrix and weight matrix obtained by the correlation calculation, respectively. <math display="inline"><semantics> <msup> <mi>V</mi> <mo>′</mo> </msup> </semantics></math> is the high-resolution feature representation of DEM-LR.</p>
Full article ">Figure 4
<p>(<b>a</b>) Inner Mongolian Plateau, (<b>b</b>) Tarim Basin, (<b>c</b>) Qinling Mountains, (<b>d</b>) North China Plain.</p>
Full article ">Figure 5
<p>DEM reconstruction visualization results: (<b>a1</b>–<b>a4</b>) is the original DEM, (<b>b1</b>–<b>b4</b>) is the reconstruction DEM.</p>
Full article ">Figure 6
<p>Comparison of DEM reconstruction visualization results of four methods in four areas.</p>
Full article ">Figure 7
<p>Histogram statistics of the DEM reconstruction, (<b>a</b>) represents area 3 and (<b>b</b>) area 4.</p>
Full article ">Figure 8
<p>A comparison of the results of DEM reconstruction using hillshade visualization for the four methods.</p>
Full article ">Figure 9
<p>Comparison of the results of the slope visualization for DEM reconstruction using four different methods in area 4.</p>
Full article ">Figure 10
<p>Comparison of the DEM reconstruction results for aspect visualization using four methods in area 4.</p>
Full article ">
20 pages, 2510 KiB  
Article
Parallel Computation of Multi-GNSS and Multi-Frequency Inter-Frequency Clock Biases and Observable-Specific Biases
by Linyang Li, Zhen Yang, Zhen Jia and Xin Li
Remote Sens. 2023, 15(7), 1953; https://doi.org/10.3390/rs15071953 - 6 Apr 2023
Cited by 1 | Viewed by 1694
Abstract
With the widespread application of GNSS, the delicate handling of biases among different systems and different frequencies is of critical importance, wherein the inter-frequency clock biases (IFCBs) and observable-specific signal biases (OSBs) should be carefully corrected. Usually, a serial approach is used to [...] Read more.
With the widespread application of GNSS, the delicate handling of biases among different systems and different frequencies is of critical importance, wherein the inter-frequency clock biases (IFCBs) and observable-specific signal biases (OSBs) should be carefully corrected. Usually, a serial approach is used to calculate these products. To accelerate the computation speed and reduce the time delay, a multicore parallel estimation strategy for IFCBs, code, and phase OSBs by utilizing task parallel library (TPL) is proposed, the parallel computations, including precise point positioning (PPP), IFCBs, and OSBs estimation, being carried out on the basis of data parallelisms and task-based asynchronous programming. Three weeks of observables from the multi-GNSS experiment campaign (MGEX) network is utilized. The result shows that the IFCB errors of GPS Block IIF and GLONASS M+ satellites are nonnegligible, in which the GLONASS M+ satellite R21 shows the largest IFCB of more than 0.60 m, while those of other systems and frequencies are marginal, and the code OSBs present excellent stability with a standard deviation (STD) of 0.10 ns for GPS and approximately 0.20 ns for other satellite systems. Besides, the phase OSBs of all systems show the stability of better than 0.10 ns, wherein the Galileo satellites show the best performance of 0.01 ns. Compared with the single-core serial computing method, the acceleration rates for IFCBs and OSBs estimation are 3.10, 5.53, 9.66, and 17.04 times higher using four, eight, sixteen, and thirty-two physical cores, respectively, through multi-core parallelized execution. Full article
(This article belongs to the Special Issue Precise Point Positioning with GPS, GLONASS, BeiDou, and Galileo II)
Show Figures

Figure 1

Figure 1
<p>The finer-grained parallel computing flowchart of undifferenced and uncombined PPP under a multicore platform. The symbol * represents the station name.</p>
Full article ">Figure 2
<p>The parallel estimation flowchart of IFCBs, code, and phase OSBs under a multicore platform. Three parallel computing methods involving inter-station, inter-satellite, and matrix operation are employed. The symbol * represents the station name.</p>
Full article ">Figure 3
<p>The distribution of 340 globally-distributed IGS MGEX tracking stations used in the experiment.</p>
Full article ">Figure 4
<p>The multi-GNSS IFCB series and maximum values on DOY 290, 2021. A different color line represents a different satellite. The orange and blue lines of GLONASS denote R21 and R09 satellites, respectively. The abbreviation G, E, R, C2, and C3 represents GPS, Galileo, GLONASS, BDS-2, and BDS-3, respectively.</p>
Full article ">Figure 5
<p>The multi-GNSS IFCB series and maximum values from DOY 290, 2021 to DOY 310, 2021. A different color line represents a different satellite.</p>
Full article ">Figure 6
<p>The multi-GNSS code OSB series from DOY 290, 2021 to DOY 310, 2021. A different color line represents a different satellite.</p>
Full article ">Figure 6 Cont.
<p>The multi-GNSS code OSB series from DOY 290, 2021 to DOY 310, 2021. A different color line represents a different satellite.</p>
Full article ">Figure 7
<p>The multi-GNSS phase OSB series on DOY 290, 2021. A different color line represents a different satellite.</p>
Full article ">Figure 8
<p>The average STDs of multi-GNSS phase OSBs from DOY 290, 2021 to DOY 310, 2021.</p>
Full article ">
22 pages, 8449 KiB  
Article
Accuracy Assessment of High-Resolution Globally Available Open-Source DEMs Using ICESat/GLAS over Mountainous Areas, A Case Study in Yunnan Province, China
by Menghua Li, Xiebing Yin, Bo-Hui Tang and Mengshi Yang
Remote Sens. 2023, 15(7), 1952; https://doi.org/10.3390/rs15071952 - 6 Apr 2023
Cited by 7 | Viewed by 2571
Abstract
The Open-Source Digital Elevation Model (DEM) is fundamental data of the geoscientific community. However, the variation of its accuracy with land cover type and topography has not been thoroughly studied. This study evaluates the accuracy of five globally covered and open-accessed DEM products [...] Read more.
The Open-Source Digital Elevation Model (DEM) is fundamental data of the geoscientific community. However, the variation of its accuracy with land cover type and topography has not been thoroughly studied. This study evaluates the accuracy of five globally covered and open-accessed DEM products (TanDEM-X90 m, SRTEM, NASADEM, ASTER GDEM, and AW3D30) in the mountain area using ICESat/GLAS data as the GCPs. The robust evaluation indicators were utilized to compare the five DEMs’ accuracy and explore the relationship between these errors and slope, aspect, landcover types, and vegetation coverage, thereby revealing the consistency differences in DEM quality under different geographical feature conditions. The Taguchi method is introduced to quantify the impact of these surface characteristics on DEM errors. The results show that the slope is the main factor affecting the accuracy of DEM products, accounting for about 90%, 81%, 85%, 83%, and 65% for TanDEM-X90, SRTM, NASADEM, ASTER GDEM, and AW3D30, respectively. TanDEM-X90 has the highest accuracy in very flat areas (slope < 2°), NASADEM and SRTM have the greatest accuracy in flat areas (2 ≤ slope < 5°), while AW3D30 accuracy is the best in other cases and shows the best consistency on slopes. This study makes a new attempt to quantify the factors affecting the accuracy of DEM, and the results can guide the selection of open-source DEMs in related geoscience research. Full article
Show Figures

Figure 1

Figure 1
<p>Study area: Yunnan Province, China. The colour changes on the graphic reflect changes in surface elevation, while the red dots indicate the ICESat satellite’s laser footprint on the ground.</p>
Full article ">Figure 2
<p>Schematic of DEMs comparison with ICESat/GLAS data.</p>
Full article ">Figure 3
<p>The distribution of the number of GCPs: (<b>a</b>) distribution on different elevations; (<b>b</b>) distribution on different slopes.</p>
Full article ">Figure 4
<p>The Q−Q diagram of the normality test of residuals. <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mrow> <mi>diff</mi> </mrow> </msub> </mrow> </semantics></math> is the height difference between the DEM products on top of the subfigures and the ICESat/GLAS data. The red line is the case of the ideal normal distribution. The blue line shows the actual situation of each DEM product.</p>
Full article ">Figure 5
<p>(<b>a</b>) Boxplot of five DEM products in the study area. (<b>b</b>) The error distributions of these five DEMs.</p>
Full article ">Figure 6
<p>The differences (i.e., bias) between five DEM products and ICESat/GLAS GCP points, shown as boxplots (with 5 degree intervals) and hex-bin scatterplots according to slope change: (<b>a</b>) TanDEM-X90, (<b>b</b>) SRTM, (<b>c</b>) NASA, (<b>d</b>) ASTER GDEM, (<b>e</b>) AW3D30.</p>
Full article ">Figure 7
<p>The differences (i.e., bias) between five DEM products and ICESat/GLAS GCP points, shown as boxplots (45 degree intervals) and hex-bin scatter plots according to aspect change: (<b>a</b>) TanDEM-X90, (<b>b</b>) SRTM, (<b>c</b>) NASA, (<b>d</b>) ASTER GDEM, and (<b>e</b>) AW3D30.</p>
Full article ">Figure 8
<p>Relationship between quality indicators ((<b>top</b>) <span class="html-italic">MAE</span> and (<b>bottom</b>) <span class="html-italic">NMAD</span>) and (<b>a</b>) slope and (<b>b</b>) aspect.</p>
Full article ">Figure 9
<p>The differences (i.e., bias) between five DEM products and ICESat/GLAS GCP points, shown boxplots and distribution histograms according to different landcover types: (<b>a</b>) Tan-DEM-X90, (<b>b</b>) SRTM, (<b>c</b>) NASA, (<b>d</b>) ASTER GDEM, and (<b>e</b>) AW3D30.</p>
Full article ">Figure 10
<p>The differences (i.e., bias) between five DEM products and ICESat GCP, shown as boxplots (with 0.1 intervals) and hex-bin scatter plots according to FVCover changes: (<b>a</b>) Tan-DEM-X, (<b>b</b>) SRTM, (<b>c</b>) NASA, (<b>d</b>) ASTER GDEM, and (<b>e</b>) AW3D30.</p>
Full article ">Figure 11
<p>Relationship between quality indicators ((<b>top</b>) <span class="html-italic">MAE</span> and (<b>bottom</b>) <span class="html-italic">NMAD</span>) and (<b>a</b>) FVCover (<b>b</b>) landcover types.</p>
Full article ">Figure 12
<p>Landcover and its relationship to slope.</p>
Full article ">Figure 13
<p>Slope-grouped boxplots of differences (i.e., bias) according to landcover types.</p>
Full article ">Figure 14
<p>Relationship between quality indicators (<span class="html-italic">MAE</span> on the top and <span class="html-italic">NMAD</span> on the bottom) and slope under different landcover types. (<b>a</b>) TanDEM-X90, (<b>b</b>) SRTM, (<b>c</b>) NASA, (<b>d</b>) ASTER GDEM, and (<b>e</b>) AW3D30.</p>
Full article ">Figure 15
<p>S/N of 16 experiments under different combinations of four factors.</p>
Full article ">Figure 16
<p>Percentage of contribution of each factor to MAE in the study area.</p>
Full article ">Figure 17
<p>(<b>a</b>) The left curve in the figure is the ICESat/GLAS echo waveform curve from the vegetation-covered flat ground, and the blue dotted line represents the actual position of the ICESat elevation point. The right side of the figure shows that the penetration of the DEM into the forest canopy is different due to the different wavelengths used by the sensors generating DEM products. The approximate depth of penetration is shown by the arrow. (<b>b</b>) The proportion of GCP points in different landcover types in the study area.</p>
Full article ">
29 pages, 16759 KiB  
Article
A Novel Hybrid Intelligent SOPDEL Model with Comprehensive Data Preprocessing for Long-Time-Series Climate Prediction
by Zeyu Zhou, Wei Tang, Mingyang Li, Wen Cao and Zhijie Yuan
Remote Sens. 2023, 15(7), 1951; https://doi.org/10.3390/rs15071951 - 6 Apr 2023
Cited by 8 | Viewed by 2573
Abstract
Long-time-series climate prediction is of great significance for mitigating disasters; promoting ecological civilization; identifying climate change patterns and preventing floods, drought and typhoons. However, the general public often struggles with the complexity and extensive temporal range of meteorological data when attempting to accurately [...] Read more.
Long-time-series climate prediction is of great significance for mitigating disasters; promoting ecological civilization; identifying climate change patterns and preventing floods, drought and typhoons. However, the general public often struggles with the complexity and extensive temporal range of meteorological data when attempting to accurately forecast climate extremes. Sequence disorder, weak robustness, low characteristics and weak interpretability are four prevalent shortcomings in predicting long-time-series data. In order to resolve these deficiencies, our study gives a novel hybrid spatiotemporal model which offers comprehensive data preprocessing techniques, focusing on data decomposition, feature extraction and dimensionality upgrading. This model provides a feasible solution to the puzzling problem of long-term climate prediction. Firstly, we put forward a Period Division Region Segmentation Property Extraction (PD-RS-PE) approach, which divides the data into a stationary series (SS) for an Extreme Learning Machine (ELM) prediction and an oscillatory series (OS) for a Long Short-term Memory (LSTM) prediction to accommodate the changing trend of data sequences. Secondly, a new type of input-output mapping mode in a three-dimensional matrix was constructed to enhance the robustness of the prediction. Thirdly, we implemented a multi-layer technique to extract features of high-speed input data based on a Deep Belief Network (DBN) and Particle Swarm Optimization (PSO) for parameter searching of a neural network, thereby enhancing the overall system’s learning ability. Consequently, by integrating all the above innovative technologies, a novel hybrid SS-OS-PSO-DBN-ELM-LSTME (SOPDEL) model with comprehensive data preprocessing was established to improve the quality of long-time-series forecasting. Five models featuring partial enhancements are discussed in this paper and three state-of-the-art classical models were utilized for comparative experiments. The results demonstrated that the majority of evaluation indices exhibit a significant optimization in the proposed model. Additionally, a relevant evaluation system showed that the quality of “Excellent Prediction” and “Good Prediction” exceeds 90%, and no data with “Bad Prediction” appear, so the accuracy of the prediction process is obviously insured. Full article
(This article belongs to the Special Issue Artificial Intelligence for Weather and Climate)
Show Figures

Figure 1

Figure 1
<p>Overall flow chart of this paper.</p>
Full article ">Figure 2
<p>Process of the proposed SOPDEL model.</p>
Full article ">Figure 3
<p>Normalized raw data for the one region studied for climate prediction.</p>
Full article ">Figure 4
<p>Novel PD−RS−PE technology.</p>
Full article ">Figure 5
<p>DBN structure.</p>
Full article ">Figure 6
<p>Improvement of input−output mapping pattern.</p>
Full article ">Figure 7
<p>PSO−DBN−ELME algorithm.</p>
Full article ">Figure 8
<p>PSO−DBN−LSTME algorithm.</p>
Full article ">Figure 9
<p>Correlation−map of HT, LT, RF, SF and SR in the study areas.</p>
Full article ">Figure 9 Cont.
<p>Correlation−map of HT, LT, RF, SF and SR in the study areas.</p>
Full article ">Figure 10
<p>Scatter plots of observed and predicted highest temperature (unit: °C) using M1−M6 models.</p>
Full article ">Figure 11
<p>Scatter plots of observed and predicted lowest temperature (unit: °C) using M1−M6 models.</p>
Full article ">Figure 12
<p>Scatter plots of observed and predicted rainfall (unit: mm) using M1−M6 models.</p>
Full article ">Figure 13
<p>Predicting effect of the test highest temperature (unit: °C).</p>
Full article ">Figure 14
<p>Predicting effect of the test lowest temperature (unit: °C).</p>
Full article ">Figure 15
<p>Predicting effect of the test rainfall (unit: mm).</p>
Full article ">Figure 16
<p>Scatter plots of the observed and predicted highest temperature (unit: °C) using M6–M9 models.</p>
Full article ">Figure 17
<p>Scatter plots of the observed and predicted lowest temperature (unit: °C) using M6−M9 models.</p>
Full article ">Figure 18
<p>Scatter plots of observed and predicted rainfall (unit: mm) using M6–M9 models.</p>
Full article ">Figure 19
<p>Predicting effect of the observed and predicted highest temperature (unit: °C) using M6–M9 models.</p>
Full article ">Figure 20
<p>Predicting effect of the observed and predicted lowest temperature (unit: °C) using M6−M9 models.</p>
Full article ">Figure 21
<p>Predicting effect of the observed and predicted rainfall (unit: mm) using M6–M9 models.</p>
Full article ">Figure 22
<p>The proportion of each quality in the predicting climate data.</p>
Full article ">
28 pages, 20227 KiB  
Article
Long-Range Lightning Interferometry Using Coherency
by Xue Bai and Martin Füllekrug
Remote Sens. 2023, 15(7), 1950; https://doi.org/10.3390/rs15071950 - 6 Apr 2023
Viewed by 1893
Abstract
Traditional lightning detection and location networks use the time of arrival (TOA) technique to locate lightning events with a single time stamp. This contribution introduces a simulation study to lay the foundation for new lightning location concepts. Here, a novel interferometric method is [...] Read more.
Traditional lightning detection and location networks use the time of arrival (TOA) technique to locate lightning events with a single time stamp. This contribution introduces a simulation study to lay the foundation for new lightning location concepts. Here, a novel interferometric method is studied which expands the data use and maps lightning events into an area by using coherency. The amplitude waveform bank, which consists of averaged waveforms classified by their propagation distances, is first used to test interferometric methods. Subsequently, the study is extended to individual lightning event waveforms. Both amplitude and phase coherency of the analytic signal are used here to further develop the interferometric method. To determine a single location for the lightning event and avoid interference between the ground wave and the first skywave, two solutions are proposed: (1) use a small receiver network and (2) apply an impulse response function to the recorded waveforms, which uses an impulse to represent the lightning occurrence. Both methods effectively remove the first skywave interference. This study potentially helps to identify the lightning ground wave without interference from skywaves with a long-range low frequency (LF) network. It is planned to expand the simulation work with data reflecting a variety of ionospheric and geographic scenarios. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Interpolated waveform bank spectrum (190–2150 km). (<b>b</b>) Sensitivity map using quality <span class="html-italic">q</span> value. The receivers are marked in red dots. The range is 35°N to 60°N for latitude and 10°W to 25°E for longitude with a pixel resolution of 0.5° × 0.5°.</p>
Full article ">Figure 2
<p>(<b>a</b>) An example of coherency waveforms calculated over randomly selected individual events with the case numbers 10, 40, 70, and 100. (<b>b</b>) The peak coherency (black), threshold coherency (blue), and ratio <span class="html-italic">R</span> (red) with different event numbers. (<b>c</b>) coherency based on different receiver pairs. The peak coherency are the lines larger than 0.4 while the threshold coherency are the lines lower than 0.4. The threshold value of 0.4 is marked with a red dashed line. (<b>d</b>) Ratio <span class="html-italic">R</span> values with different selections of receiver locations. In both (<b>c</b>,<b>d</b>), the upper figure shows the coherency at Bath (red), Toulouse (green), Orleans (black), and Rustrel (blue). The middle figure shows the coherency for the location pairs Bath–Toulouse (red), Bath–Orleans (green), Bath–Rustrel (black), Toulouse–Orleans (blue), Toulouse–Rustrel (sky blue), and Orleans–Rustrel (purple). The bottom figure shows the coherency for the location sets Toulouse–Orleans–Rustrel (red), Bath–Orleans–Rustrel (green), Bath–Toulouse–Rustrel (black), and Bath–Toulouse–Orleans (blue).</p>
Full article ">Figure 3
<p>(<b>a</b>) Distance distribution for events recorded from different locations. (<b>b</b>) Noise coherency with different event numbers. This figure shows theoretical coherency noise values (blue), the coherency of noises recorded at Bath (yellow), Toulouse (purple), Orleans (green), and Rustrel (sky, blue), and event threshold coherency (red).</p>
Full article ">Figure 4
<p>(<b>a</b>) Simulated lightning locations (red open circle and red dot) and receiver distribution in Europe (all the dots except two red dots). The red open circle is the simulated lightning location using the averaged waveforms from the amplitude waveform bank, and the red dot is the simulated lightning location of 15°E and 42°S. Yellow dots are the receiver used in the simulation with averaged waveforms from the amplitude waveform bank; black dots are the 10 receivers used in a small receiver network (<span class="html-italic">N</span> = 10); the black dots and green dots are the 20 receivers used in small receiver network (<span class="html-italic">N</span> = 20). (<b>b</b>) The calculated coherency map by using the amplitude waveforms from the amplitude waveform bank. The coherency map has a latitude and longitude range of 0.5° × 0.5°, with a precision of 0.001° × 0.001°.</p>
Full article ">Figure 5
<p>Amplitude waveforms from the amplitude waveform bank used to simulate the coherency map.</p>
Full article ">Figure 6
<p>(<b>a</b>) Coherency waveforms used in the simulation with 105 receiver network. Calculated coherency waveform of the selected individual events (<b>upper</b>) and the coherency waveform of 1150 km from the coherency waveform bank (<b>bottom</b>). The ground wave maxima are marked with red dots, and the skywave maxima are marked with red open circles. (<b>b</b>) The coherency waveform calculated by the filtered waveforms. The maximum is marked with a red dot.</p>
Full article ">Figure 7
<p>The coherency maps with 105 receiver network with a simulated lightning location of 15°E and 42°S. Each sub-figure (<b>a</b>–<b>l</b>) is the coherency map at a different time.</p>
Full article ">Figure 8
<p>(<b>a</b>) Amplitude waveforms used in the simulation with 105 receiver network. The middle figure is the scaled amplitude waveform; the bottom figure is the ratioed amplitude waveform; and the upper figure is the amplitude waveform of 1150 km from the amplitude waveform bank. The ground wave maxima are marked with red dots, and the skywave maxima are marked with red open circles. (<b>b</b>) The amplitude waveform calculated by the filtered waveforms. The maximum is marked with a red dot.</p>
Full article ">Figure 9
<p>The scaled amplitude maps with 105 receiver network with a simulated lightning location of 15°E and 42°S. Each sub-figure (<b>a</b>–<b>l</b>) is the amplitude map of a different time frame.</p>
Full article ">Figure 10
<p>(<b>a</b>) Coherency waveforms used in the simulation with a small network (<span class="html-italic">N</span> = 10). (<b>b</b>) Coherency waveforms used in the simulation with a small network (<span class="html-italic">N</span> = 20). For both Figures (<b>a</b>,<b>b</b>), the upper figure is the calculated coherency waveform and the bottom figure is the coherency waveform of 270 km and 410 km, respectively, from the coherency waveform bank. In all the figures, the ground wave maxima are marked with red dots.</p>
Full article ">Figure 11
<p>The coherency maps with a small receiver network with a simulated lightning location of 15°E and 42°S (<span class="html-italic">N</span> = 10). Each sub-figure (<b>a</b>–<b>i</b>) is the coherency map at a different time.</p>
Full article ">Figure 12
<p>The coherency maps with a small receiver network with a simulated lightning location of 15°E and 42°S (<span class="html-italic">N</span> = 20). Each sub-figure (<b>a</b>–<b>i</b>) is the coherency map at a different time.</p>
Full article ">Figure 13
<p>(<b>a</b>) Amplitude waveforms used in the simulation with a small network (<span class="html-italic">N</span> = 10). (<b>b</b>) Amplitude waveforms used in the simulation with a small network (<span class="html-italic">N</span> = 20). For both figures (<b>a</b>,<b>b</b>), the upper figures are the averaged waveforms from the amplitude waveform bank (270 km and 410 km); the middle figures are the scaled amplitude waveforms; and the bottom figures are the ratioed amplitude waveforms. In all the figures, the ground wave maxima are marked with red dots.</p>
Full article ">Figure 14
<p>The scaled amplitude maps with a small receiver network with a simulated lightning location of 15°E and 42°S (<span class="html-italic">N</span> = 10).</p>
Full article ">Figure 15
<p>The scaled amplitude maps with a small receiver network with a simulated lightning location of 15°E and 42°S (<span class="html-italic">N</span> = 20).</p>
Full article ">Figure 16
<p>The coherency maps using the impulse response function filtered waveforms with a simulated lightning location of 15°E and 42°S. Each sub-figure (<b>a</b>–<b>i</b>) is the coherency map at a different time.</p>
Full article ">Figure 17
<p>The amplitude maps using the impulse response function filtered waveforms with a simulated lightning location of 15°E and 42°S. Each sub-figure (<b>a</b>–<b>i</b>) is the amplitude map at a different time.</p>
Full article ">
26 pages, 15778 KiB  
Article
A Comparative Assessment of Multi-Source Generation of Digital Elevation Models for Fluvial Landscapes Characterization and Monitoring
by Paweł Sudra, Luca Demarchi, Grzegorz Wierzbicki and Jarosław Chormański
Remote Sens. 2023, 15(7), 1949; https://doi.org/10.3390/rs15071949 - 6 Apr 2023
Cited by 5 | Viewed by 2598
Abstract
Imaging and measuring the Earth’s relief with sensors mounted upon unmanned aerial vehicles is an increasingly frequently used and promising method of remote sensing. In the context of fluvial geomorphology and its applications, e.g., landform mapping or flood modelling, the reliable representation of [...] Read more.
Imaging and measuring the Earth’s relief with sensors mounted upon unmanned aerial vehicles is an increasingly frequently used and promising method of remote sensing. In the context of fluvial geomorphology and its applications, e.g., landform mapping or flood modelling, the reliable representation of the land surface on digital elevation models is crucial. The main objective of the study was to assess and compare the accuracy of state-of-the-art remote sensing technologies in generating DEMs for riverscape characterization and fluvial monitoring applications. In particular, we were interested in DAP and LiDAR techniques comparison, and UAV applicability. We carried out field surveys, i.e., GNSS-RTK measurements, UAV and aircraft flights, on islands and sandbars within a nature reserve on a braided section of the Vistula River downstream from the city of Warsaw, Poland. We then processed the data into DSMs and DTMs based on four sources: ULS (laser scanning from UAV), UAV-DAP (digital aerial photogrammetry), ALS (airborne laser scanning), and satellite Pléiades imagery processed with DAP. The magnitudes of errors are represented by the cross-reference of values generated on DEMs with GNSS-RTK measurements. Results are presented for exposed sediment bars, riverine islands covered by low vegetation and shrubs, or covered by riparian forest. While the average absolute height error of the laser scanning DTMs oscillates around 8–11 cm for most surfaces, photogrammetric DTMs from UAV and satellite data gave errors averaging more than 30 cm. Airborne and UAV LiDAR measurements brought almost the perfect match. We showed that the UAV-based LiDAR sensors prove to be useful for geomorphological mapping, especially for geomorphic analysis of the river channel at a large scale, because they reach similar accuracies to ALS and better than DAP-based image processing. Full article
(This article belongs to the Special Issue Remote Sensing of Riparian Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Location of the “Ławice Kiełpińskie” natural reserve in the context of Warsaw urban agglomeration. Background map: OpenStreetMap. Data sources: EEA/GDOŚ, GUGiK, Solon et al., 2018 [<a href="#B38-remotesensing-15-01949" class="html-bibr">38</a>], own elaboration.</p>
Full article ">Figure 2
<p>Study area within the “Ławice Kiełpińskie” reserve. Oblique photo taken from a UAV in July 2019 showing a panoramic view upstream, towards Warsaw: northern part located in the Jabłonna commune. Exposed sandbars and islands with varying levels of vegetation can be seen.</p>
Full article ">Figure 3
<p>Study area. Islands within the “Ławice Kiełpińskie” (“Kielpinskie Shoals”) natural reserve. Overlayed are acquisition sections of the airborne and UAV missions, and the distribution of ground control points from GNSS-RTK measurements. Background orthophotomap: Google Satellite.</p>
Full article ">Figure 4
<p>Changes in the water level at the Warsaw “Bulwary” water gauge within the period of UAV and airborne data acquisitions. Hydrological data provided by the Institute of Meteorology and Water Management (IMGW), courtesy of T. Lewicki.</p>
Full article ">Figure 5
<p>Flowchart of data processing as regards data obtained from the UAV platform—laser scanning point cloud processing and photogrammetric procedures, used to generate DEMs.</p>
Full article ">Figure 6
<p>Example of the UAV-LiDAR point cloud for the Southern Island, representing the heterogenous landscape characterized by different land covers of the studied islands. Visualization made in LiDAR 360 viewer.</p>
Full article ">Figure 7
<p>Digital surface models’ (DSMs) coverage and heights comparison (<b>A</b> = UAV-LS, <b>B</b> = UAV-DAP, <b>C</b> = ALS, <b>D</b> = PLEIADES) on the example of Northern Island of the “Kiełpińskie Shoals”. Resolution of the models is always 10 cm, except for the Pléiades model (50 cm).</p>
Full article ">Figure 8
<p>Accuracy of DTMs generated from data from different platforms and sensors: unmanned aerial vehicle (UAV-LiDAR and UAV-DAP), aircraft (ALS), and Pléiades satellite (DAP–RGB images).</p>
Full article ">Figure 9
<p>Accuracy of DTMs generated from data from different platforms and sensors, grouped by different land cover classes: unmanned aerial vehicle (LiDAR and DAP techniques), aircraft (ALS–LiDAR), and Pléiades satellite (DAP–RGB images).</p>
Full article ">Figure 10
<p>Accuracy of DTMs generated from LiDAR data obtained from UAV platform and from the aircraft (ALS) in two spectral channels separately and combined. The comparison is grouped by the different land cover classes analysed.</p>
Full article ">Figure 11
<p>Scatterplots of height values generated on DTM according to different data sources (from top: UAV-LiDAR, ALS—airborne LiDAR, UAV—RGB photogrammetry, satellite images from Pléiades), and according to land cover classes ((<b>a</b>)—uncovered sandbars, (<b>b</b>)—shrub vegetation, (<b>c</b>)—riparian forest). The measurement values are in meters, WGS84 ellipsoidal (geodetic) heights.</p>
Full article ">Figure 12
<p>DTM visualization of the Northern Island (central-western section) with height error values at the GCP locations. Digital elevation model based on laser scanning from UAV (ULS). GCPs shown correspond to two vegetation and geomorphological classes: (1) shrub vegetation on the riverine islands and transforming higher level of sandbars, (2) uncovered sandy deposits on sandbars. Background orthophotomap: Google Satellite (with partial transparency).</p>
Full article ">Figure 13
<p>(<b>a</b>) Northern Island cross-section used to make the profile graph, image <b>A</b>—ALS, image <b>B</b>—UAV-DAP, image <b>C</b>—UAV-LS (ULS), (<b>b</b>) Northern Island—profile lines comparing three digital terrain models (DTMs). The measurement values are in metres, WGS84 ellipsoidal (geodetic heights).</p>
Full article ">Figure 14
<p>(<b>a</b>) Southern Island cross-section used to make the profile graph, image <b>A</b>—ALS, image <b>B</b>—UAV-LS, (<b>b</b>) Southern Island—profile lines comparing two LIDAR-based digital terrain models (DTMs). The measurement values are in metres, WGS84 ellipsoidal (geodetic heights).</p>
Full article ">Figure 15
<p>(<b>a</b>) Northern Island and Vistula riverbed cross-section used to make the profile graph, (<b>b</b>) Northern Island—comparison of four digital surface models (DSMs) based on different data sources (UAV-LS, ALS, UAV-DAP, PLEIADES). The measurement values are in metres, WGS84 ellipsoidal (geodetic heights).</p>
Full article ">
24 pages, 35825 KiB  
Article
The Respondence of Wave on Sea Surface Temperature in the Context of Global Change
by Ru Yao, Weizeng Shao, Mengyu Hao, Juncheng Zuo and Song Hu
Remote Sens. 2023, 15(7), 1948; https://doi.org/10.3390/rs15071948 - 6 Apr 2023
Cited by 10 | Viewed by 2103
Abstract
Several aspects of global climate change, e.g., the rise of sea level and water temperature anomalies, suggest the advantages of studying wave distributions. In this study, WAVEWATCH-III (WW3) (version 6.07), which is a well-known numerical wave model, was employed for simulating waves over [...] Read more.
Several aspects of global climate change, e.g., the rise of sea level and water temperature anomalies, suggest the advantages of studying wave distributions. In this study, WAVEWATCH-III (WW3) (version 6.07), which is a well-known numerical wave model, was employed for simulating waves over global seas from 1993–2020. The European Centre for Medium-Range Weather Forecasts (ECMWF), Copernicus Marine Environment Monitoring Service (CMEMS), current and sea level were used as the forcing fields in the WW3 model. The validation of modelling simulations against the measurements from the National Data Buoy Center (NDBC) buoys and Haiyang-2B (HY-2B) altimeter yielded a root mean square error (RMSE) of 0.49 m and 0.63 m, with a correlation (COR) of 0.89 and 0.90, respectively. The terms calculated by WW3-simulated waves, i.e., breaking waves, nonbreaking waves, radiation stress, and Stokes drift, were included in the water temperature simulation by a numerical circulation model named the Stony Brook Parallel Ocean Model (sbPOM). The water temperature was simulated in 2005–2015 using the high-quality Simple Ocean Data Assimilation (SODA) data. The validation of sbPOM-simulated results against the measurements obtained from the Array for Real-time Geostrophic Oceanography (Argo) buoys yielded a RMSE of 1.12 °C and a COR of 0.99. By the seasonal variation, the interrelation of the currents, sea level anomaly, and significant wave heights (SWHs) were strong in the Indian Ocean. In the strong current areas, the distribution of the sea level was consistent with the SWHs. The monthly variation of SWHs, currents, sea surface elevation, and sea level anomalies revealed that the upward trends of SWHs and sea level anomalies were consistent from 1993–2015 over the global ocean. In the Indian Ocean, the SWHs were obviously influenced by the SST and sea surface wind stress. The rise of wind stress intensity and sea level enlarges the growth of waves, and the wave-induced terms strengthen the heat exchange at the air–sea layer. It was assumed that the SST oscillation had a negative response to the SWHs in the global ocean from 2005–2015. This feedback indicates that the growth of waves could slow down the amplitude of water warming. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The daily-average Copernicus Marine Environment Monitoring Service (CMEMS) currents map on 1 January 2020; (<b>b</b>) the daily-average CMEMS sea level map on 1 January 2020; and (<b>c</b>) the European Centre for Medium-Range Weather Forecasts (ECMWF) winds map at 00:00 UTC on 1 January 2020. The red lines in (<b>a</b>) are the dividing lines of the Pacific Ocean, Atlantic Ocean, and Indian Ocean.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) The daily-average Copernicus Marine Environment Monitoring Service (CMEMS) currents map on 1 January 2020; (<b>b</b>) the daily-average CMEMS sea level map on 1 January 2020; and (<b>c</b>) the European Centre for Medium-Range Weather Forecasts (ECMWF) winds map at 00:00 UTC on 1 January 2020. The red lines in (<b>a</b>) are the dividing lines of the Pacific Ocean, Atlantic Ocean, and Indian Ocean.</p>
Full article ">Figure 2
<p>The geographic locations of the National Data Buoy Center (NDBC) buoys and Array for Real-time Geostrophic Oceanography (Argo) buoys in the Pacific Ocean. The green points represent the NDBC buoys and the red points represent the Argo buoys, in which the numbers beside the buoys represent the IDs.</p>
Full article ">Figure 3
<p>The SWH measurements from the Haiyang-2B (HY-2B) altimeter from 00:00 to 04:00 UTC on 15 January 2020. The background was the SWH simulated from the WW3 model at 00:00 UTC on 15 January 2020.</p>
Full article ">Figure 4
<p>(<b>a</b>) The Simple Ocean Data Assimilation (SODA) practical salinity map; (<b>b</b>) the SODA potential temperature map in January 2015; (<b>c</b>) the National Centers for Environmental Prediction (NCEP) total heat flux; (<b>d</b>) the global bathymetric topography map.</p>
Full article ">Figure 5
<p>(<b>a</b>) Comparison of the SWHs simulated using the WW3 model with NDBC buoys for a 0.4 m bin. (<b>b</b>) Comparison of the WW3-simulated SWHs with HY-2B measurements for a 0.4 m bin. (<b>c</b>) Comparison of the sea surface temperature simulated using the sbPOM model with Argos for a 1.8 °C bin between −3 °C and 33 °C. Note that the error bars represent the standard deviation of each bin for the matchups.</p>
Full article ">Figure 6
<p>The seasonal-averaged significant wave height (SWH) (m) at the global ocean (60°S–60°N, 180°W–180°E) calculated by the WW3 model. (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 7
<p>The seasonal-averaged CMEMS current speed (m/s) at the global ocean (60°S–60°N, 180°W–180°E). (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 8
<p>The relationship between the monthly-averaged SWH and CMEMS currents. (<b>a</b>) the Global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean; (<b>d</b>) the Indian Ocean. The red and black solid lines represent the current speed and the SWH, respectively.</p>
Full article ">Figure 9
<p>The seasonal-averaged CMEMS sea surface elevation at global ocean (60°S–60°N, 180°W–180°E). (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 10
<p>The seasonal-averaged CMEMS sea level anomaly at global ocean (60°S–60°N, 180°W–180°E). (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 11
<p>The relationship between monthly-averaged SWH and CMEMS sea level. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean; (<b>d</b>) the Indian Ocean. The red and black solid lines represent the water surface elevation and the SWH, respectively.</p>
Full article ">Figure 12
<p>The relationship between the monthly-averaged SWH and CMEMS sea level anomaly. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean; (<b>d</b>) the Indian Ocean. The red and black solid lines represent the water surface elevation and the sea level anomaly, respectively.</p>
Full article ">Figure 13
<p>The relationship between the monthly-averaged SWH anomaly and CMEMS sea level anomaly. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean; (<b>d</b>) the Indian Ocean. The red and black solid lines represent the SWH anomaly and the sea level anomaly, respectively.</p>
Full article ">Figure 14
<p>The seasonal-averaged sbPOM-simulated sea surface temperature (°C) at the global ocean (60°S–60°N, 180°W–180°E). (<b>a</b>) spring; (<b>b</b>) summer; (<b>c</b>) autumn; (<b>d</b>) winter.</p>
Full article ">Figure 15
<p>The relationship between monthly-averaged SWH and sbPOM sea surface temperature. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean; (<b>d</b>) the Indian Ocean. The red and black solid lines represent the sea surface temperature and the SWH, respectively.</p>
Full article ">Figure 16
<p>The relationship between monthly-averaged sbPOM sea surface temperature and sea surface wind stress. (<b>a</b>) The global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean, and (<b>d</b>) the Indian Ocean. The red and black solid lines represent the sea surface wind stress and the sea surface temperature, respectively.</p>
Full article ">Figure 17
<p>The relationship between monthly-averaged SWH and sea surface wind stress. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean, and (<b>d</b>) the Indian Ocean. The red and black solid lines represent the sea surface wind stress and the SWH, respectively.</p>
Full article ">Figure 17 Cont.
<p>The relationship between monthly-averaged SWH and sea surface wind stress. (<b>a</b>) the global ocean; (<b>b</b>) the Pacific Ocean; (<b>c</b>) the Atlantic Ocean, and (<b>d</b>) the Indian Ocean. The red and black solid lines represent the sea surface wind stress and the SWH, respectively.</p>
Full article ">Figure 18
<p>The linear regression analysis of the WW3-simulated SWH and (<b>a</b>) currents from CMEMS, (<b>b</b>) sea level from CMEMS, (<b>c</b>) sbPOM-simulated sea surface temperature, (<b>d</b>) CMEMS sea level anomaly, and (<b>e</b>) sea surface wind stress in the Pacific Ocean. The linear regression analysis of the WW3-simulated SWH anomaly and CMEMS sea level anomaly (<b>f</b>) in the Pacific Ocean. The black solid line represents the linear fitted line. The two red dotted lines represent the 95% prediction intervals.</p>
Full article ">Figure 19
<p>The linear regression analysis of WW3-simulated SWH and (<b>a</b>) currents from CMEMS, (<b>b</b>) sea level from CMEMS, (<b>c</b>) sbPOM-simulated sea surface temperature, (<b>d</b>) CMEMS sea level anomaly, and (<b>e</b>) sea surface wind stress in the Atlantic Ocean. The linear regression analysis of the WW3-simulated SWH anomaly and the CMEMS sea level anomaly (<b>f</b>) in the Atlantic Ocean. The black solid line represents the linear fitted line. The two red dotted lines represent the 95% prediction intervals.</p>
Full article ">Figure 20
<p>The linear regression analysis of WW3-simulated SWH and (<b>a</b>) currents from CMEMS, (<b>b</b>) sea level from CMEMS, (<b>c</b>) sbPOM-simulated sea surface temperature, (<b>d</b>) the CMEMS sea level anomaly, and (<b>e</b>) the sea surface wind stress in the Indian Ocean. The linear regression analysis of WW3-simulated SWH anomaly and CMEMS sea level anomaly (<b>f</b>) in the Indian Ocean. The black solid line represents the linear fitted line. The two red dotted lines represent the 95% prediction intervals.</p>
Full article ">
22 pages, 12963 KiB  
Article
Multi-Resolution Population Mapping Based on a Stepwise Downscaling Approach Using Multisource Data
by Yan Jin, Rui Liu, Haoyu Fan, Pengdu Li, Yaojie Liu and Yan Jia
Remote Sens. 2023, 15(7), 1947; https://doi.org/10.3390/rs15071947 - 6 Apr 2023
Cited by 1 | Viewed by 2335
Abstract
The distribution of the population is an essential aspect of addressing social, economic, and environmental problems. Gridded population data can provide more detailed information than census data, and multisource data from remote sensing and geographic information systems have been widely used for population [...] Read more.
The distribution of the population is an essential aspect of addressing social, economic, and environmental problems. Gridded population data can provide more detailed information than census data, and multisource data from remote sensing and geographic information systems have been widely used for population estimation studies. However, due to spatial heterogeneity, the population has different distribution characteristics and variation patterns at different scales, while the relationships between multiple variables also vary with scale. This article presents a stepwise downscaling approach in that the random forest regression kriging technique is used to downscale census data to multi-resolution gridded population datasets. Using Nanjing, China, as the experimental case, population distribution maps were generated at 100 m, 500 m, and 1 km spatial resolution, and compared with the other three downscaling methods and three population products. The results demonstrated the produced gridded population maps by the proposed approach have higher accuracy and more accurate details of population distribution with the smallest mean absolute error (MAE) and root mean squared error (RMSE) values of 1.590 and 2.189 ten thousand people (over 40% reduction). The artificial land and road data are the two most important indicators of population distribution for the regional random forest modeling in Nanjing. Our proposed method can be a valuable tool for population mapping and has the potential to monitor sustainable development goals. Full article
(This article belongs to the Section Remote Sensing and Geo-Spatial Science)
Show Figures

Figure 1

Figure 1
<p>Nanjing Study area.</p>
Full article ">Figure 2
<p>Census data at administrative street zones of Nanjing in (<b>a</b>) 2010 and (<b>b</b>) 2020. The absent street census in 2020 were supplemented by the corresponding district census, shown in purple to lighter purples.</p>
Full article ">Figure 3
<p>Flowchart of the multi-resolution population mapping methodology.</p>
Full article ">Figure 4
<p>The feature importance of the top-10 indicators in three steps: (<b>a</b>) 1 km regression models; (<b>b</b>) 500 m regression models; (<b>c</b>) 100 m regression models.</p>
Full article ">Figure 5
<p>The spatial distribution of downscaled population in 2010 (<b>a</b>–<b>d</b>) at 1 km resolution, (<b>e</b>–<b>h</b>) at 500 m resolution, and (<b>i</b>–<b>l</b>) at 100 m resolution by using MLR, RF, MLRK, and RFRK methods, respectively.</p>
Full article ">Figure 6
<p>The spatial distribution of downscaled population in 2020 (<b>a</b>–<b>d</b>) at 1 km resolution, (<b>e</b>–<b>h</b>) at 500 m resolution, and (<b>i</b>–<b>l</b>) at 100 m resolution by using MLR, RF, MLRK, and RFRK methods, respectively.</p>
Full article ">Figure 7
<p>Four comparison metrics for different downscaling models in (<b>a</b>) both years and (<b>b</b>) single year. The average values were calculated in both years.</p>
Full article ">Figure 8
<p>The absolute error at administrative street zones between census data and aggregated population predictions of 1 km and 100 m resolutions in 2010 and 2020: (<b>a</b>–<b>d</b>) WorldPop products; (<b>e</b>–<b>h</b>) RFRK-based downscaled results. The absent street zones in 2020 were filled with aggregated WorldPop data or aggregated RFRK-based predictions (marked as RFK).</p>
Full article ">Figure 9
<p>The feature importance of the top-10 indicators in three steps by using XGBoost: (<b>a</b>) 1 km regression models; (<b>b</b>) 500 m regression models; (<b>c</b>) 100 m regression models.</p>
Full article ">Figure 10
<p>The LCRPGR valued in Nanjing in 2010–2020. The spatial distribution at different resolutions: (<b>a</b>) administrative district zones by using census data; (<b>b</b>) administrative street zones by using 100 m downscaled population based on RFRK; (<b>c</b>) 1 km resolution by using 1 km downscaled population based on RFRK.</p>
Full article ">
18 pages, 6835 KiB  
Article
The Development of a Low-Cost Hydrophone for Passive Acoustic Monitoring of Dolphin’s Vocalizations
by Rocco De Marco, Francesco Di Nardo, Alessandro Lucchetti, Massimo Virgili, Andrea Petetta, Daniel Li Veli, Laura Screpanti, Veronica Bartolucci and David Scaradozzi
Remote Sens. 2023, 15(7), 1946; https://doi.org/10.3390/rs15071946 - 6 Apr 2023
Cited by 4 | Viewed by 3651
Abstract
Passive acoustics are widely used to monitor the presence of dolphins in the marine environment. This study aims to introduce a low-cost and homemade approach for assembling a complete underwater microphone (i.e., the hydrophone), employing cheap and easy to obtain components. The hydrophone [...] Read more.
Passive acoustics are widely used to monitor the presence of dolphins in the marine environment. This study aims to introduce a low-cost and homemade approach for assembling a complete underwater microphone (i.e., the hydrophone), employing cheap and easy to obtain components. The hydrophone was assembled with two piezo disks connected in a balanced configuration and encased in a plastic container filled with plastic foam. The hydrophone’s performance was validated by direct comparison with the commercially available AS-1 hydrophone (Aquarian Hydrophones, Anacortes, U.S.) on different underwater acoustic signals: artificial acoustic signals (ramp and multitone signals) and various dolphin vocalizations (whistle, echolocation clicks, and burst pulse signals). The sensitivity of the device’s performance to changes in the emission source position was also tested. The results of the validation procedure on both artificial signals and real dolphin vocalizations showed that the significant cost savings associated with cheap technology had a minimal effect on the recording device’s performance within the frequency range of 0–35 kHz. At this stage of experimentation, the global cost of the hydrophone could be estimated at a few euros, making it extremely price competitive when compared to more expensive commercially available models. In the future, this effective and low-cost technology would allow for continuous monitoring of the presence of free-ranging dolphins, significantly lowering the total cost of autonomous monitoring systems. This would permit broadening the monitored areas and creating a network of recorders, thus improving the acquisition of data. Full article
(This article belongs to the Special Issue Remote Sensing and Other Geomatics Techniques for Marine Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The proposed CoPiDi hydrophone.</p>
Full article ">Figure 2
<p>Schematic representation of the proposed homemade preamplifier.</p>
Full article ">Figure 3
<p>(<b>a</b>) Connection scheme of the devices used in the experimental trials. The reference AS-1 hydrophone is closely coupled with a PA4 preamp and the CoPiDi hydrophone has been connected to two different preamplifiers near to the audio device. (<b>b</b>) Experimental arrangement in the dolphin pool. (<b>c</b>) Multitone and ramp signals trials configuration.</p>
Full article ">Figure 4
<p>Portion of the multitone source signal produced by the STM DDD pinger.</p>
Full article ">Figure 5
<p>(<b>a</b>) The haul scheme used in experimental dolphin trials: the CoPiDi and reference AS-1 hydrophones were placed at a distance of 100 cm and at the same depth of 120 cm. (<b>b</b>) A map of the Oltremare marine park. The star indicates the location of the hydrophones.</p>
Full article ">Figure 6
<p>Example of a single ramp signal spectrogram (in green) and PSD analysis comparison. (<b>a</b>) Response of the Reference Hydrophone (<b>b</b>) Response of CoPiDi Hydrophone.</p>
Full article ">Figure 7
<p>Comparative analysis of PSD in the frequency domain between the proposed hydrophone and the reference hydrophone. Four different orientations and two different preamplifiers were tested (<b>a</b>) Signal source at 0°. (<b>b</b>) Signal source at 90°. (<b>c</b>) Signal source at 180°. (<b>d</b>) Hydrophone in orizontal position.</p>
Full article ">Figure 8
<p>Comparison of the results of the multitone signal analysis between the CoPiDi (panel <b>a</b>) and the reference (panel <b>b</b>) hydrophones. The blue points represent PSD peak values detected by the two hydrophones on a single 30-s recording.</p>
Full article ">Figure 9
<p>Hydrophone performance in a representative 5-s segment using the proposed low-cost approach (spectrogram in panel <b>a</b> and click detection in panel <b>b</b>) as compared to the reference approach (spectrogram in panel <b>c</b> and click detection in panel <b>d</b>). Detected clicks are highlighted with a circle. SNR is signal-to-noise ratio.</p>
Full article ">Figure 10
<p>Example of a comparison between CoPiDi hydrophone (<b>a</b>) and the reference hydrophone (<b>b</b>) by visual inspection during PAM analysis between the waveforms of click sounds (red box), feeding buzzes (yellow box), and whistles (blue box) detected by means of the two different approaches.</p>
Full article ">
16 pages, 13281 KiB  
Article
Convolutional Neural Network Maps Plant Communities in Semi-Natural Grasslands Using Multispectral Unmanned Aerial Vehicle Imagery
by Maren Pöttker, Kathrin Kiehl, Thomas Jarmer and Dieter Trautz
Remote Sens. 2023, 15(7), 1945; https://doi.org/10.3390/rs15071945 - 6 Apr 2023
Cited by 6 | Viewed by 3505
Abstract
Semi-natural grasslands (SNGs) are an essential part of European cultural landscapes. They are an important habitat for many animal and plant species and offer a variety of ecological functions. Diverse plant communities have evolved over time depending on environmental and management factors in [...] Read more.
Semi-natural grasslands (SNGs) are an essential part of European cultural landscapes. They are an important habitat for many animal and plant species and offer a variety of ecological functions. Diverse plant communities have evolved over time depending on environmental and management factors in grasslands. These different plant communities offer multiple ecosystem services and also have an effect on the forage value of fodder for domestic livestock. However, with increasing intensification in agriculture and the loss of SNGs, the biodiversity of grasslands continues to decline. In this paper, we present a method to spatially classify plant communities in grasslands in order to identify and map plant communities and weed species that occur in a semi-natural meadow. For this, high-resolution multispectral remote sensing data were captured by an unmanned aerial vehicle (UAV) in regular intervals and classified by a convolutional neural network (CNN). As the study area, a heterogeneous semi-natural hay meadow with first- and second-growth vegetation was chosen. Botanical relevés of fixed plots were used as ground truth and independent test data. Accuracies up to 88% on these independent test data were achieved, showing the great potential of the usage of CNNs for plant community mapping in high-resolution UAV data for ecological and agricultural applications. Full article
(This article belongs to the Special Issue Crops and Vegetation Monitoring with Remote/Proximal Sensing)
Show Figures

Figure 1

Figure 1
<p>Location of the study site in Germany (<b>top left</b>) and the district of Osnabrück (<b>bottom left</b>). Orthomosaic and grassland vegetation of one plot of 06/08/2021 (<b>right</b>).</p>
Full article ">Figure 2
<p>Schematic workflow of preprocessing, training, validation, and classification.</p>
Full article ">Figure 3
<p>Scatter plots of the samples of the dependent (○) and independent (+) test data in blue vs. green and red vs. infrared. Colors are used as follows: <b>grey</b>: <span class="html-italic">Rumex obtusifolius</span> plants, <b>blue</b>: <span class="html-italic">Lolium perenne</span>-community, <b>red</b>: <span class="html-italic">Alopecurus pratensis</span>-community, <b>green</b>: <span class="html-italic">Bromus Hordeaceus</span>-community.</p>
Full article ">Figure 4
<p>Subsets of the classification results of the mono- and multitemporal model and orthomosaics in RGB-color of <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mn>1</mn> </msub> <msub> <mi>T</mi> <mn>3</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>G</mi> <mn>2</mn> </msub> <msub> <mi>T</mi> <mn>3</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">
16 pages, 2343 KiB  
Article
Decline of Late Spring and Summer Snow Cover in the Scottish Highlands from 1984 to 2022: A Landsat Time Series
by Benedict D. Spracklen and Dominick V. Spracklen
Remote Sens. 2023, 15(7), 1944; https://doi.org/10.3390/rs15071944 - 6 Apr 2023
Cited by 2 | Viewed by 2047
Abstract
Late spring and summer snow cover, the remnants of winter and early spring snowfall, not only possess an intrinsic importance for montane flora and fauna, but also act as a sensitive indicator for climate change. The variability and potential trends in late spring [...] Read more.
Late spring and summer snow cover, the remnants of winter and early spring snowfall, not only possess an intrinsic importance for montane flora and fauna, but also act as a sensitive indicator for climate change. The variability and potential trends in late spring and summer (snowmelt season) snow cover in mountain regions are often poorly documented. May to mid-September Landsat imagery from 1984 to 2022 was used to quantify changes in the snow-covered area of upland regions in the Scottish Highlands. There was substantial annual variability in the area of May to mid-September snow cover combined with a significant decline over the 39-year study period (p = 0.02). Long-term climate data used to show variability in May to mid-September snow cover was positively related to winter snowfall and negatively related to winter and April temperatures. The results from a long-running field survey counting the number of snow patches that survive until the following winter were used to check the veracity of the study. Further, accuracy was estimated through comparison with higher resolution Sentinel-2 imagery, giving a user and producer accuracy rate of 99.8% and 87%, respectively. Projected future warming will further diminish this scarce, valuable habitat, along with its associated plant communities, thus threatening the biodiversity and scenic value of the Scottish Highlands. Full article
Show Figures

Figure 1

Figure 1
<p>Map of location of study areas with background shading showing elevation (m).</p>
Full article ">Figure 2
<p>Number of Landsat images used to determine snow cover for each year in the study.</p>
Full article ">Figure 3
<p>Overall workflow to determine snow cover value (SV) from Landsat imagery. LT is the abbreviation for Landsat; Eqn for Equation.</p>
Full article ">Figure 4
<p>Curves of snow-covered pixels versus date for the nine sites in the study area. The red line shows the best-fit line for each site. Note the logarithmic scale of the y-axis.</p>
Full article ">Figure 5
<p>Mean annual snow cover from 1984 to 2022. Error bars show standard error. Positive values of SV indicate more snow than normal, negative values indicate less than normal.</p>
Full article ">
12 pages, 817 KiB  
Communication
Design of Extensible Structured Interferometric Array Utilizing the “Coarray” Concept
by Qiang Wang, Cong Xue, Shurui Zhang, Renli Zhang and Weixing Sheng
Remote Sens. 2023, 15(7), 1943; https://doi.org/10.3390/rs15071943 - 5 Apr 2023
Cited by 1 | Viewed by 1787
Abstract
The optimum placement of receiving telescope antennas is a central topic for designing radio interferometric arrays, and this determines the performance of the obtained information. A variety of arrays are designed for different purposes, and they perform poorly in scalability. In this paper, [...] Read more.
The optimum placement of receiving telescope antennas is a central topic for designing radio interferometric arrays, and this determines the performance of the obtained information. A variety of arrays are designed for different purposes, and they perform poorly in scalability. In this paper, we consider a subclass of structured sparse arrays, namely nested arrays, and examine the important role of “coarray” in interferometric synthesis imaging, which is utilized to design nested array configurations for a complete uniform Fourier plane coverage in both supersynthesis and instantaneous modes. Both nested arrays and the theory of the coarray have rich research achievements, and we apply them to astronomy to design arrays with good scalability and imaging performance. Simulated celestial source image retrieval performance validates the effectiveness of nested interferometric arrays. Full article
(This article belongs to the Special Issue SAR, Interferometry and Polarimetry Applications in Geoscience)
Show Figures

Figure 1

Figure 1
<p>Configuration of linear nested array and corresponding difference coarray.</p>
Full article ">Figure 2
<p>Configuration of Y-shaped nested array and corresponding difference coarray.</p>
Full article ">Figure 3
<p>Configurations of cross-product and 2D nested arrays.</p>
Full article ">Figure 4
<p>The u-v mask of linear nested array and corresponding dirty beam.</p>
Full article ">Figure 5
<p>The u-v mask of Y-shaped nested array and corresponding dirty beam.</p>
Full article ">Figure 6
<p>The instantaneous u-v mask and corresponding dirty beam.</p>
Full article ">Figure 7
<p>The variations of FR and PSL for three different extended Y-shaped arrays.</p>
Full article ">Figure 8
<p>True 2D Gaussian source and retrieved images by different nested arrays. (<b>a</b>) True 2D Gaussian source for Y-shaped nested array with 27 antennas. (<b>b</b>) Linear nested array. (<b>c</b>) Y-shaped nested array. (<b>d</b>) 2D nested array.</p>
Full article ">Figure 9
<p>Retrieved images by extended linear nested arrays.</p>
Full article ">Figure 10
<p>Retrieved images by extended Y-shaped nested arrays.</p>
Full article ">
23 pages, 9006 KiB  
Article
Terrestrial Laser Scanning for Non-Destructive Estimation of Aboveground Biomass in Short-Rotation Poplar Coppices
by María Menéndez-Miguélez, Guillermo Madrigal, Hortensia Sixto, Nerea Oliveira and Rafael Calama
Remote Sens. 2023, 15(7), 1942; https://doi.org/10.3390/rs15071942 - 5 Apr 2023
Cited by 6 | Viewed by 2170
Abstract
Poplar plantations in high-density and short-rotation coppices (SRC) are a suitable way for the fast production of wood that can be transformed into bioproducts or bioenergy. Optimal management of these coppices requires accurate assessment of the total standing biomass. However, traditional field inventory [...] Read more.
Poplar plantations in high-density and short-rotation coppices (SRC) are a suitable way for the fast production of wood that can be transformed into bioproducts or bioenergy. Optimal management of these coppices requires accurate assessment of the total standing biomass. However, traditional field inventory is a challenging task, given the existence of multiple shoots, the difficulty of identifying terminal shoots, and the extreme high density. As an alternative, in this work, we propose to develop individual stool and plot biomass models using metrics derived from terrestrial laser scanning (TLS) as predictors. To this aim, we used data from a SRC poplar plantation, including nine plots and 154 individual stools. Every plot was scanned from different positions, and individual stools were felled, weighed, and dried to compute aboveground biomass (AGB). Individual stools were segmented from the cloud point, and different TLS metrics at stool and plot level were derived following processes of bounding box, slicing, and voxelization. These metrics were then used, either alone or combined with field-measured metrics, to fit biomass models. Our results indicate that at individual-stool level, the biomass models combining TLS metrics and easy to measure in field metrics (stool diameter) perform similarly to the traditional allometric models based on field inventories, while at plot scales, TLS-derived models show superiority over traditional models. Our proposed methodology permits accurate and non-destructive estimates of the biomass fixed in SRC plantations. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of TLS data acquired from the FARO Focus3D M70 scanner in one of the study plots. Stools represented by green circles; scan positions represented by blue asterisks; spheres represented by red triangles.</p>
Full article ">Figure 2
<p>Detail of one of the study plots (<b>left</b>), and the target spheres and the FARO Focus<sup>3D</sup> M70 scanner during the scanning process (<b>right</b>).</p>
Full article ">Figure 3
<p>Example of TLS data acquired from the FARO Focus<sup>3D</sup> M70 scanner in one of the study plots: overhead view (<b>left</b>) and side view (<b>right</b>) showing detail from the studied stools.</p>
Full article ">Figure 4
<p>Representation of the three algorithms used to derive stool volume from TLS data. From left to right and top to bottom: picture of the scanned stool; bounding vox; slices; and voxels (2 cm, 5 cm, 10 cm, and 25 cm).</p>
Full article ">Figure 5
<p>Concordance between observed and predicted values of individual stool biomass for the TLS model (<b>a</b>), Field-inventory model (<b>b</b>), and combined model (<b>c</b>). Straight lines indicate the 1:1.</p>
Full article ">Figure 6
<p>Effect of clone over the relation between individual stool biomass and the height derived from TLS slicing (<span class="html-italic">H_slice</span>). Shaded indicate 95% confidence intervals.</p>
Full article ">Figure 7
<p>Concordance between observed and predicted values of total biomass for the TLS model (<b>a</b>), Field-inventory model (<b>b</b>), and combined model (<b>c</b>). Straight lines indicate the 1:1.</p>
Full article ">
18 pages, 7948 KiB  
Article
Monitoring Glacier Lake Outburst Flood (GLOF) of Lake Merzbacher Using Dense Chinese High-Resolution Satellite Images
by Changjun Gu, Suju Li, Ming Liu, Kailong Hu and Ping Wang
Remote Sens. 2023, 15(7), 1941; https://doi.org/10.3390/rs15071941 - 5 Apr 2023
Cited by 8 | Viewed by 4439
Abstract
Establishing an effective real-time monitoring and early warning system for glacier lake outburst floods (GLOFs) requires a full understanding of their occurrence mechanism. However, the harsh conditions and hard-to-reach locations of these glacial lakes limit detailed fieldwork, making satellite imagery a critical tool [...] Read more.
Establishing an effective real-time monitoring and early warning system for glacier lake outburst floods (GLOFs) requires a full understanding of their occurrence mechanism. However, the harsh conditions and hard-to-reach locations of these glacial lakes limit detailed fieldwork, making satellite imagery a critical tool for monitoring. Lake Mercbacher, an ice-dammed lake in the central Tian Shan mountain range, poses a significant threat downstream due to its relatively high frequency of outbursts. In this study, we first monitored the daily changes in the lake area before the 2022 Lake Mercbacher outburst. Additionally, based on historical satellite images from 2014 to 2021, we calculated the maximum lake area (MLA) and its changes before the outburst. Furthermore, we extracted the proportion of floating ice and water area during the period. The results show that the lake area of Lake Mercbacher would first increase at a relatively low speed (0.01 km2/day) for about one month, followed by a relatively high-speed increase (0.04 km2/day) until reaching the maximum, which would last for about twenty days. Then, the lake area would decrease slowly until the outburst, which would last five days and is significant for early warning. Moreover, the floating ice and water proportion provides more information about the outburst signals. In 2022, we found that the floating ice area increased rapidly during the early warning stage, especially one day before the outburst, accounting for about 50% of the total lake area. Historical evidence indicates that the MLA shows a decreasing trend, and combining it with the outburst date and climate data, we found that the outburst date shows an obvious advance trend (6 days per decade) since 1902, caused by climate warming. Earlier melting results in an earlier outburst. This study provides essential references for monitoring Lake Mercbacher GLOFs and building an effective early warning system. Full article
Show Figures

Figure 1

Figure 1
<p>The location of lake Merzbacher.</p>
Full article ">Figure 2
<p>Satellite image (<b>a</b>) and image after calculating the NDWI (<b>b</b>).</p>
Full article ">Figure 3
<p>Maximum lake area (MLA) change before outburst since 2014. (<b>a</b>) MLA extent in each year and overlay on the image obtained on 1 August 2014; (<b>b</b>) MLA changes from 2014 to 2022; (<b>c</b>–<b>j</b>) MLA overlay on images obtained from 2015 to 2022.</p>
Full article ">Figure 4
<p>Lake extent change in 2022 before and after the outburst.</p>
Full article ">Figure 5
<p>Centroid changes of lake area in historical and 2022 before the outburst.</p>
Full article ">Figure 6
<p>Water area and Ice area of Lake Merzbacher before the outburst from 2014 to 2022 (<b>a</b>); Water area and Ice area daily changes before the outburst (<b>b</b>).</p>
Full article ">Figure 7
<p>Date of Lake Merzbacher outburst since 1902. (<b>a</b>) Lake Merzbacher outburst Frequency in different months since 1902; (<b>b</b>) Lake Merzbacher outburst date since 1902; (<b>c</b>) The relationship between the outburst date and the hottest date since 1902.</p>
Full article ">Figure 8
<p>Detailed records of outburst date from Lake Merzbacher from 1902 to 2022 (<b>a</b>). Temperature changes in Lake Merzbacher from 1981 to 2022 (<b>b</b>).</p>
Full article ">Figure 9
<p>Satellite image of Lake Merzbacher captured in 4 May 2022.</p>
Full article ">
18 pages, 18279 KiB  
Article
On Surface Waves Generated by Extra-Tropical Cyclones—Part I: Multi-Satellite Measurements
by Vahid Cheshm Siyahi, Vladimir Kudryavtsev, Maria Yurovskaya, Fabrice Collard and Bertrand Chapron
Remote Sens. 2023, 15(7), 1940; https://doi.org/10.3390/rs15071940 - 5 Apr 2023
Cited by 4 | Viewed by 1878
Abstract
Surface waves generated by Extra-Tropical Cyclones (ETCs) can significantly affect shipping, fishing, offshore oil and gas production, and other marine activities. This paper presents the results of a satellite data-based investigation of wind waves generated by two North Atlantic ETCs. These ETCs were [...] Read more.
Surface waves generated by Extra-Tropical Cyclones (ETCs) can significantly affect shipping, fishing, offshore oil and gas production, and other marine activities. This paper presents the results of a satellite data-based investigation of wind waves generated by two North Atlantic ETCs. These ETCs were fast-moving systems, inhibiting resonance (synchronism) between the group velocity of the generated waves and the ETC translation velocity. In these cases, wave generation begins when the front boundary of the storm appears at a given ocean location point. Since developing waves are slow, they move backward relative to the storm, grow in time, and then leave the ETC stormy area through the rear sector. Multi-satellite observations confirm such a paradigm, revealing that the storm regions are filled with young developing wind waves, the most developed in the rear-right sector. As observed, the energy of these waves grew in time during the ETC life span. It is demonstrated that the extended-fetch concept (inherent for Tropical Cyclones) does not apply to ETC. Instead, by analogy, the concept of extended-duration wave growth is more relevant. Satellite observations confirmed the validity of duration-laws for waves generated by ETCs, and demonstrated that extended-fetch solutioncan be valid at time scales exceeding the lifespan of considered ETCs. Full article
Show Figures

Figure 1

Figure 1
<p>Mean sea level pressure, MSLP, maps with wind speed contours in ms<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>. White contours are the storm area of ETCs with the gray circle as their center.</p>
Full article ">Figure 2
<p>Time evolution of ETC#1 (<b>a</b>) and ETC#2 (<b>b</b>) parameters: (red thick line) maximum wind speed, <math display="inline"><semantics> <msub> <mi>u</mi> <mi>m</mi> </msub> </semantics></math>; (red dashed line) translation velocity, <math display="inline"><semantics> <msub> <mi>V</mi> <mi>t</mi> </msub> </semantics></math>; (blue thick line) radius of maximum wind speed, <math display="inline"><semantics> <msub> <mi>R</mi> <mi>m</mi> </msub> </semantics></math>; (blue dashed line) critical fetch, <math display="inline"><semantics> <msub> <mover> <mi>L</mi> <mo>¯</mo> </mover> <mrow> <mi>c</mi> <mi>r</mi> </mrow> </msub> </semantics></math> from (<a href="#FD1-remotesensing-15-01940" class="html-disp-formula">1</a>).</p>
Full article ">Figure 3
<p>Altimetry tracks passing through the Northern Atlantic during the period of 11 February 2020 00:00 UTC–16 February 2020 00:00 UTC, after applying ice and land masks. The black line shows the trajectory of the ETC#1 and the gray line ETC#2. The colour of the altimeter tracks indicates measured SWH in meters.</p>
Full article ">Figure 4
<p>Measured <math display="inline"><semantics> <msub> <mi>λ</mi> <mi>p</mi> </msub> </semantics></math> (colored boxes) and <math display="inline"><semantics> <msub> <mi>φ</mi> <mi>p</mi> </msub> </semantics></math> after eliminating 180° ambiguity (black arrows) in “wave boxes” on two sides of the CFOSAT-SWIM track in Northern Atlantic during the period of 11 February 2020 00:00 UTC —16 February 2020 00:00 UTC, after applying ice and land masks. The green and cyan colored lines and arrows show the trajectory and direction of the ETC#1 and ETC#2, respectively.</p>
Full article ">Figure 5
<p>Wind speed and SWH along the AltiKa, CryoSat-2, Sentinel-3, and CFOSAT-SWIM nadir tracks at the different moments of ETC#1 lifetime. Columns from left to right, (1st column): NCEP/CFSv2 wind field and position of the track; (2nd column): along-track wind speed from (black line) measurements and (red lines) NCEP/CFSv2; (3rd column): profile of measured SWH (black line) and SWH of fully-developed waves for local wind speed (red line); (4th column): zoom on the ETC#1 storm (inner) area (shown in the first column by red contour). The red arrows in the fourth column indicate the ETC’s direction, and the radius of dashed circles changes from 200 km at a 200 km interval. Vertical shaded areas in columns 2 and 3 indicate the parts of the tracks that fell into the ETC’s storm area, shown in the first column by red contour. The position of <math display="inline"><semantics> <msub> <mi>u</mi> <mi>m</mi> </msub> </semantics></math> is indicated by a black asterisk.</p>
Full article ">Figure 6
<p>The same as in <a href="#remotesensing-15-01940-f005" class="html-fig">Figure 5</a>, but for ETC#2 and altimeters AltiKa, CryoSat-2, Sentinel-3, Jason-3, and CFOSAT-SWIM nadir tracks.</p>
Full article ">Figure 6 Cont.
<p>The same as in <a href="#remotesensing-15-01940-f005" class="html-fig">Figure 5</a>, but for ETC#2 and altimeters AltiKa, CryoSat-2, Sentinel-3, Jason-3, and CFOSAT-SWIM nadir tracks.</p>
Full article ">Figure 7
<p>Two−dimensional scatter plot of significant wave height SWH as a function of wind speed, colored with points density. The black curve shows the <math display="inline"><semantics> <mrow> <msub> <mi>H</mi> <msub> <mi>s</mi> <mrow> <mi>f</mi> <mi>d</mi> </mrow> </msub> </msub> <mo>=</mo> <mn>0.21</mn> <mo> </mo> <msubsup> <mi>u</mi> <mrow> <mn>10</mn> </mrow> <mn>2</mn> </msubsup> <mo>/</mo> <mi>g</mi> </mrow> </semantics></math> [<a href="#B35-remotesensing-15-01940" class="html-bibr">35</a>].</p>
Full article ">Figure 8
<p>Along−track profile of (<b>a</b>,<b>e</b>,<b>i</b>) SWHs, <math display="inline"><semantics> <msub> <mi>H</mi> <mi>s</mi> </msub> </semantics></math>; (<b>b</b>,<b>f</b>,<b>j</b>) spectral peak wavelength, <math display="inline"><semantics> <msub> <mi>λ</mi> <mi>p</mi> </msub> </semantics></math>; (<b>c</b>,<b>g</b>,<b>k</b>) inverse wave age,<math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mrow> <mo>|</mo> <mo>|</mo> </mrow> </msub> <mo>=</mo> <mrow> <mo>(</mo> <mi>u</mi> <mo>/</mo> <msub> <mi>c</mi> <mi>p</mi> </msub> <mo>)</mo> </mrow> <mo form="prefix">cos</mo> <mrow> <mo>(</mo> <msub> <mi>φ</mi> <mi>p</mi> </msub> <mo>−</mo> <msub> <mi>φ</mi> <mi>w</mi> </msub> <mo>)</mo> </mrow> </mrow> </semantics></math>; and (<b>d</b>,<b>h</b>,<b>l</b>) wave steepens, <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <msub> <mi>k</mi> <mi>p</mi> </msub> <msub> <mi>H</mi> <mi>s</mi> </msub> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>k</mi> <mi>p</mi> </msub> <mo>=</mo> <mn>2</mn> <mi>π</mi> <mo>/</mo> <msub> <mi>λ</mi> <mi>p</mi> </msub> </mrow> </semantics></math>. Blue circles and curves show the measurements, and the orange curves represent the values associated with fully developed waves. The average time of the SWIM passage over the North Atlantic Ocean is shown on the right side of each row. The SWIM nadir tracks can be found in 2nd, 8th and 10th rows of <a href="#remotesensing-15-01940-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 9
<p>Along−track wind speed, SWH, and <math display="inline"><semantics> <mi>α</mi> </semantics></math> inside the cyclone in a coordinate system associated with the coordinates of ETCs considering their heading aligned in <span class="html-italic">x</span>-direction, respectively, illustrated in (<b>a</b>–<b>c</b>) for ETC#1 and (<b>d</b>–<b>f</b>) for ETC#2.</p>
Full article ">Figure 10
<p>Time evolution of maximum SWH values for ETC#1 (<b>a</b>) ETC#2 (<b>b</b>). The filled red circles are related to the SWH of wind waves (<math display="inline"><semantics> <mrow> <mi>α</mi> <mo>&gt;</mo> <mn>0.85</mn> </mrow> </semantics></math>), and the open red circles—to the SWH of swell. The black curves show the SWH of fully developed waves calculated for local <math display="inline"><semantics> <msub> <mi>u</mi> <mi>m</mi> </msub> </semantics></math>.</p>
Full article ">Figure 11
<p>Dimensionless energy, <math display="inline"><semantics> <mover accent="true"> <mi>e</mi> <mo>˜</mo> </mover> </semantics></math>, versus dimensionless time, <math display="inline"><semantics> <mover accent="true"> <mi>t</mi> <mo>˜</mo> </mover> </semantics></math>, for ETC#1 (<b>a</b>) and ETC#2 (<b>b</b>): red circles are observations, and green lines are duration laws (<a href="#FD6-remotesensing-15-01940" class="html-disp-formula">6</a>). Dashed-black and dashed-blue lines indicate the dimensionless time interval <math display="inline"><semantics> <msub> <mi>t</mi> <mn>0</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> </semantics></math> defined by (<a href="#FD7-remotesensing-15-01940" class="html-disp-formula">7</a>) and (<a href="#FD8-remotesensing-15-01940" class="html-disp-formula">8</a>), and corresponding dimensionless energy predicted by duration laws (<a href="#FD6-remotesensing-15-01940" class="html-disp-formula">6</a>).</p>
Full article ">
32 pages, 2644 KiB  
Review
A Comprehensive Review of Geospatial Technology Applications in Earthquake Preparedness, Emergency Management, and Damage Assessment
by Mahyat Shafapourtehrany, Maryna Batur, Farzin Shabani, Biswajeet Pradhan, Bahareh Kalantar and Haluk Özener
Remote Sens. 2023, 15(7), 1939; https://doi.org/10.3390/rs15071939 - 5 Apr 2023
Cited by 12 | Viewed by 9490
Abstract
The level of destruction caused by an earthquake depends on a variety of factors, such as magnitude, duration, intensity, time of occurrence, and underlying geological features, which may be mitigated and reduced by the level of preparedness of risk management measures. Geospatial technologies [...] Read more.
The level of destruction caused by an earthquake depends on a variety of factors, such as magnitude, duration, intensity, time of occurrence, and underlying geological features, which may be mitigated and reduced by the level of preparedness of risk management measures. Geospatial technologies offer a means by which earthquake occurrence can be predicted or foreshadowed; managed in terms of levels of preparation related to land use planning; availability of emergency shelters, medical resources, and food supplies; and assessment of damage and remedial priorities. This literature review paper surveys the geospatial technologies employed in earthquake research and disaster management. The objectives of this review paper are to assess: (1) the role of the range of geospatial data types; (2) the application of geospatial technologies to the stages of an earthquake; (3) the geospatial techniques used in earthquake hazard, vulnerability, and risk analysis; and (4) to discuss the role of geospatial techniques in earthquakes and related disasters. The review covers past, current, and potential earthquake-related applications of geospatial technology, together with the challenges that limit the extent of usefulness and effectiveness. While the focus is mainly on geospatial technology applied to earthquake research and management in practice, it also has validity as a framework for natural disaster risk assessments, emergency management, mitigation, and remediation, in general. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart representing information flow in each section.</p>
Full article ">Figure 2
<p>Global earthquake-related studies using remote sensing, published as full research papers in peer-reviewed journals and conference proceedings. Statistical information was collected using keyword searches in the academic databases Scopus and Web of Science for articles published between 2010 and 2021.</p>
Full article ">Figure 3
<p>Timeline of earthquake-related studies using remote sensing from the top ten countries with the most seismic activity between 2010 and 2021. The <span class="html-italic">Y</span>-axis shows the number of RS-related papers and <span class="html-italic">X</span>-axis indicates years of publication.</p>
Full article ">
35 pages, 1464 KiB  
Review
Overview of the Application of Remote Sensing in Effective Monitoring of Water Quality Parameters
by Godson Ebenezer Adjovu, Haroon Stephen, David James and Sajjad Ahmad
Remote Sens. 2023, 15(7), 1938; https://doi.org/10.3390/rs15071938 - 4 Apr 2023
Cited by 51 | Viewed by 13452
Abstract
This study provides an overview of the techniques, shortcomings, and strengths of remote sensing (RS) applications in the effective retrieval and monitoring of water quality parameters (WQPs) such as chlorophyll-a concentration, turbidity, total suspended solids, colored dissolved organic matter, total dissolved solids among [...] Read more.
This study provides an overview of the techniques, shortcomings, and strengths of remote sensing (RS) applications in the effective retrieval and monitoring of water quality parameters (WQPs) such as chlorophyll-a concentration, turbidity, total suspended solids, colored dissolved organic matter, total dissolved solids among others. To be effectively retrieved by RS, these WQPs are categorized as optically active or inactive based on their influence on the optical characteristics measured by RS sensors. RS applications offer the opportunity for decisionmakers to quantify and monitor WQPs on a spatiotemporal scale effectively. The use of RS for water quality monitoring has been explored in many studies using empirical, analytical, semi-empirical, and machine-learning algorithms. RS spectral signatures have been applied for the estimation of WQPs using two categories of RS, namely, microwave and optical sensors. Optical RS, which has been heavily applied in the estimation of WQPs, is further grouped as spaceborne and airborne sensors based on the platform they are on board. The choice of a particular sensor to be used in any RS application depends on various factors including cost, and spatial, spectral, and temporal resolutions of the images. Some of the known satellite sensors used in the literature and reviewed in this paper include the Multispectral Instrument aboard Sentinel-2A/B, Moderate Resolution Imaging Spectroradiometer, Landsat Thematic Mapper, Enhanced Thematic Mapper, and Operational Land Imager. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A proposed schematic framework to be utilized for monitoring and assessment of Water Quality using Remote Sensing Applications.</p>
Full article ">
27 pages, 42556 KiB  
Article
Space Target Material Identification Based on Graph Convolutional Neural Network
by Na Li, Chengeng Gong, Huijie Zhao and Yun Ma
Remote Sens. 2023, 15(7), 1937; https://doi.org/10.3390/rs15071937 - 4 Apr 2023
Cited by 4 | Viewed by 2664
Abstract
Under complex illumination conditions, the spectral data distributions of a given material appear inconsistent in the hyperspectral images of the space target, making it difficult to achieve accurate material identification using only spectral features and local spatial features. Aiming at this problem, a [...] Read more.
Under complex illumination conditions, the spectral data distributions of a given material appear inconsistent in the hyperspectral images of the space target, making it difficult to achieve accurate material identification using only spectral features and local spatial features. Aiming at this problem, a material identification method based on an improved graph convolutional neural network is proposed. Superpixel segmentation is conducted on the hyperspectral images to build the multiscale joint topological graph of the space target global structure. Based on this, topological graphs containing the global spatial features and spectral features of each pixel are generated, and the pixel neighborhoods containing the local spatial features and spectral features are collected to form material identification datasets that include both of these. Then, the graph convolutional neural network (GCN) and the three-dimensional convolutional neural network (3-D CNN) are combined into one model using strategies of addition, element-wise multiplication, or concatenation, and the model is trained by the datasets to fuse and learn the three features. For the simulated data and the measured data, the overall accuracy of the proposed method can be kept at 85–90%, and their kappa coefficients remain around 0.8. This proves that the proposed method can improve the material identification performance under complex illumination conditions with high accuracy and strong robustness. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of the space target attitude and illumination.</p>
Full article ">Figure 2
<p>Schematic of the Fresnel effect. When the incidence angle of light on the surface is larger, the specular reflection component is stronger, and the diffuse reflection component is weaker.</p>
Full article ">Figure 3
<p>Local coordinate system of the space target model.</p>
Full article ">Figure 4
<p>Normalized materials’ spectra in the hyperspectral image of the space target model.</p>
Full article ">Figure 5
<p>Normalized spectra of gold mylar in the images with different illumination conditions: (<b>a</b>) the first image; (<b>b</b>) the second image.</p>
Full article ">Figure 6
<p>An overview of the proposed method.</p>
Full article ">Figure 7
<p>Flowchart for building the topological graph of the space target global structure.</p>
Full article ">Figure 8
<p>Examples of the simulated data: (<b>a</b>) example 1; (<b>b</b>) example 2.</p>
Full article ">Figure 9
<p>Examples of normalized materials’ spectra in the simulated data.</p>
Full article ">Figure 10
<p>Laboratory measurement scene.</p>
Full article ">Figure 11
<p>Spectra of the incandescent lamp and the tungsten halogen lamp.</p>
Full article ">Figure 12
<p>Examples of the measured data: (<b>a</b>) example 1; (<b>b</b>) example 2.</p>
Full article ">Figure 13
<p>Examples of normalized materials’ spectra in the measured data.</p>
Full article ">Figure 14
<p>Examples of material identification results using the simulated data: (<b>a</b>) example 1; (<b>b</b>) example 2.</p>
Full article ">Figure 15
<p>F1-measure chart of the simulated data.</p>
Full article ">Figure 16
<p>Identification performance chart of the simulated data (OA-AA-Kappa).</p>
Full article ">Figure 17
<p>Examples of material identification results using the measured data (<b>a</b>) example 1; (<b>b</b>) example 2.</p>
Full article ">Figure 18
<p>F1-measure chart of the measured data.</p>
Full article ">Figure 19
<p>Identification performance chart of the measured data (OA-AA-Kappa).</p>
Full article ">Figure 20
<p>Distributions of grayscale values of different materials in different incidence directions and spatial resolutions: (<b>a</b>) high spatial resolution, negative X-axis; (<b>b</b>) high spatial resolution, positive Y-axis; (<b>c</b>) high spatial resolution, negative Z-axis; (<b>d</b>) low spatial resolution, positive Y-axis.</p>
Full article ">Figure 21
<p>Spectral data distributions of high spatial resolution and negative X-axis incidence: (<b>a</b>) left and right solar cells; (<b>b</b>) gold mylar on different surfaces.</p>
Full article ">Figure 22
<p>Spectral data distributions of high spatial resolution and positive Y-axis incidence: (<b>a</b>) left and right solar cells; (<b>b</b>) gold mylar on different surfaces.</p>
Full article ">Figure 23
<p>Spectral data distributions of high spatial resolution and negative Z-axis incidence; (<b>a</b>) left and right solar cells; (<b>b</b>) gold mylar on different surfaces.</p>
Full article ">Figure 24
<p>Spectral data distributions of low spatial resolution and positive Y-axis incidence: (<b>a</b>) left and right solar cells; (<b>b</b>) gold mylar on different surfaces.</p>
Full article ">Figure 25
<p>Increase the differences between target features and background features through contrastive learning.</p>
Full article ">Figure 26
<p>Schematic of staring hyperspectral images with decreasing imaging distance.</p>
Full article ">
30 pages, 19069 KiB  
Article
Insights into the Magmatic Feeding System of the 2021 Eruption at Cumbre Vieja (La Palma, Canary Islands) Inferred from Gravity Data Modeling
by Fuensanta G. Montesinos, Sergio Sainz-Maza, David Gómez-Ortiz, José Arnoso, Isabel Blanco-Montenegro, Maite Benavent, Emilio Vélez, Nieves Sánchez and Tomás Martín-Crespo
Remote Sens. 2023, 15(7), 1936; https://doi.org/10.3390/rs15071936 - 4 Apr 2023
Cited by 11 | Viewed by 3659
Abstract
This study used spatiotemporal land gravity data to investigate the 2021 eruption that occurred in the Cumbre Vieja volcano (La Palma, Canary Islands). First, we produced a density model by inverting the local gravity field using data collected in July 2005 and July [...] Read more.
This study used spatiotemporal land gravity data to investigate the 2021 eruption that occurred in the Cumbre Vieja volcano (La Palma, Canary Islands). First, we produced a density model by inverting the local gravity field using data collected in July 2005 and July 2021. This model revealed a low-density body beneath the western flank of the volcano that explains a highly fractured and altered structure related to the active hydrothermal system. Then, we retrieved changes in gravity and GNSS vertical displacements from repeated measurements made in a local network before (July 2021) and after (January 2022) the eruption. After correcting the vertical surface displacements, the gravity changes produced by mass variation during the eruptive process were used to build a forward model of the magmatic feeding system consisting of dikes and sills based on an initial model defined by the paths of the earthquake hypocenters preceding the eruption. Our study provides a final model of the magma plumbing system, which establishes a spatiotemporal framework tracing the path of magma ascent from the crust–mantle boundary to the surface from 11–19 September 2021, where the shallowest magma path was strongly influenced by the low-density body identified in the inversion process. Full article
Show Figures

Figure 1

Figure 1
<p>Simplified geological map of the island of La Palma, including historical eruptions and the last eruption in 2021 at Cumbre Vieja. UTM coordinates are in km, in zone 28. Black squares denote the villages of Puerto Naos (PN), La Bombilla (LB) and Jedey (JE).</p>
Full article ">Figure 2
<p>(<b>a</b>) Variation in the baseline length calculated for GNSS stations LP01, LPAL and MAZO situated in La Palma and station BALN in El Hierro. (<b>b</b>) Variation in the local east, north and upward coordinates for GNSS stations LP01, LPAL and MAZO in La Palma. The gray shaded areas in (<b>a</b>,<b>b</b>) indicate the period of the seismic swarm prior to the eruption onset (red line). <span class="html-italic">Y</span>-axis units are shifted for easier viewing. (<b>c</b>) Maps of the La Palma (the black area indicates the 2021 lava flows) and El Hierro islands, showing the distribution of GNSS stations (red dots).</p>
Full article ">Figure 3
<p>Distribution of the gravity stations of the different surveys used in this study. Marine gravity values are from the U.S. Geological Survey (USGS) [<a href="#B36-remotesensing-15-01936" class="html-bibr">36</a>]. The rectangular area shows an expanded image of Cumbre Vieja, where the names of absolute gravity stations (blue uppercase) and stations of the gravity network (black lowercase) are indicated. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 4
<p>Bouguer anomaly map calculated from the gravity values observed before the 2021 eruption (<a href="#remotesensing-15-01936-f003" class="html-fig">Figure 3</a>). The zone occupied by lava flows during the 2021 eruption is shaded in black, and the eruptive centers are indicated by white triangles. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 5
<p>(<b>a</b>) Regional component of the gravity field estimated by covariance analysis. The horizontal distribution of deep volcanotectonic (VT) earthquakes is included. (<b>b</b>) Corresponding local gravity map calculated by removing the regional component from the observed gravity map without (center)/with (right) the horizontal distribution of shallow volcanotectonic events. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 6
<p>(<b>a</b>) Height differences observed between January 2022 and July 2021 as calculated through GNSS data processing observed in the gravity network. (<b>b</b>) Free-air correction for the gravity stations obtained from the height data (<b>a</b>). (<b>c</b>) Gravity attraction due to lava flows and the new volcanic edifices of the 2021 eruption. (<b>d</b>) Values of gravity differences calculated between January 2022 and July 2021. (<b>e</b>) Same results as (<b>d</b>) but considering the free-air correction and gravitational attraction of the new topography. Note that the contour maps shown in (<b>d</b>,<b>e</b>) are influenced by the distribution of a few gravity stations, showing predicted values where there are no data. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 7
<p>Model of the density contrasts obtained from the inversion of the local gravity map (<a href="#remotesensing-15-01936-f005" class="html-fig">Figure 5</a>b) of the Cumbre Vieja area (La Palma). Density contrasts range between −200 and 300 kg/m<sup>3</sup>. We show several horizontal sections at several depths and in three vertical sections (at the bottom of the figure) following the S–N, W–E and NW–SE directions, which are highlighted in the horizontal section at 0 km. The hypocenters of the volcanotectonic earthquakes that occurred from July to December 2021 are included. Red triangles indicate emission centers, the red polygon shows the area occupied by the lava flows during the 2021 eruption and green squares denote the villages of Puerto Naos (PN), La Bombilla (LB) and Jedey (JE). UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 8
<p>(<b>a</b>) Plane view of the spatiotemporal distribution of volcanotectonic earthquake events (m<sub>bLg</sub> &gt; 2) from 11–19 September 2021 and syneruptive intermediate-depth events. From this information, we produced an initial model of the magma plumbing system of the 2021 eruption consisting of different dikes (1I, 3I and 5I) and sills (2I, 4I and 6I). (<b>b</b>) Three-dimensional sketch of the model shown in (<b>a</b>). The values next to the labels correspond to the thickness of the sills and the width of the dikes. Details on the geometric properties of the elements of the model can be found in the text and in <a href="#remotesensing-15-01936-t001" class="html-table">Table 1</a>. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 9
<p>(<b>a</b>) Map of the gravity response due to the initial model of the magmatic feeding system (black contour lines) (<a href="#remotesensing-15-01936-t002" class="html-table">Table 2</a>), which should fit the gravity differences observed between January 2022 and July 2021 (filled colored contour map). (<b>b</b>) The sills and dikes were approximated by prisms (black lines and red labels in (<b>c</b>)) according to the locations of the volcanotectonic earthquakes. (<b>c</b>) Map showing the differences between the observed and calculated gravity values obtained by the interpolation of the data measured at the gravity stations. The maximum value (dark red spotted area) shows a gravity increase between surveys not justified by the gravity attraction of the sills and dikes of this initial model. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 10
<p>(<b>a</b>) Final model of dikes and sills (blue lines) inferred from forward gravity modeling. The gravity response of the model is shown as contour lines. Different dipping and right prisms (<a href="#remotesensing-15-01936-t003" class="html-table">Table 3</a>) were used to define the final model that best fits the increasing gravity values (filled contour map). (<b>b</b>) Comparison with <a href="#remotesensing-15-01936-f009" class="html-fig">Figure 9</a> to help identify the differences between the initial (black lines) and final (blue lines) models. (<b>c</b>) Map of the difference between the observed and calculated gravity values obtained from the values measured in the gravity network through interpolation.he color scale is the same as that used in <a href="#remotesensing-15-01936-f009" class="html-fig">Figure 9</a>. UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 11
<p>Comparison between gravity values (<b>left</b>) and residuals (observed gravity minus calculated gravity) (<b>right</b>) corresponding to the initial and final models. The observed gravity differences between January 2022 and July 2021 are displayed in black (<b>left</b>). Error bars indicate the errors estimated from the observed gravity values and free-air correction for each station. The statistical parameters are included in the table (<b>right</b>). The distribution of the gravity stations is indicated in <a href="#remotesensing-15-01936-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 12
<p>(<b>a</b>) Vertical displacement computed with Coulomb 3.4 software based on the modeled magmatic intrusions (<a href="#remotesensing-15-01936-t003" class="html-table">Table 3</a>). (<b>b</b>) Height variations (in cm) between January 2022 and July 2021 calculated from the processed GNSS observations obtained from the gravity network during the respective surveys, displayed as brown contour lines over the vertical displacement shown in (<b>a</b>). UTM coordinates are in m, in zone 28.</p>
Full article ">Figure 13
<p>Distribution of the density contrast structures calculated from the gravity data inversion with the representation of the model of sills and dikes of the magma plumbing system (<a href="#remotesensing-15-01936-t003" class="html-table">Table 3</a>) during successive time periods before and until the onset eruption. The formation of sills (2F, 4F, 6F and 8F) at different horizontal boundaries and the path followed by dikes (1F, 3F, 5F and 7F) running through low-density structures support the hypothesis of the proposed model. The white arrows indicate the magma path. The distribution of volcanotectonic earthquakes shows a good correlation with the proposed model. The vertical scale is given in km.</p>
Full article ">
17 pages, 4983 KiB  
Article
Dynamics of Forest Vegetation in an Urban Agglomeration Based on Landsat Remote Sensing Data for the Period 1990–2022: A Case Study
by Elena Petrovna Yankovich, Ksenia Stanislavovna Yankovich and Nikolay Viktorovich Baranovskiy
Remote Sens. 2023, 15(7), 1935; https://doi.org/10.3390/rs15071935 - 4 Apr 2023
Cited by 4 | Viewed by 2027
Abstract
In recent years, the vegetation cover in urban agglomerations has been changing very rapidly due to technogenic influence. Satellite images play a huge role in studying the dynamics of forest vegetation. Special programs are used to process satellite images. The purpose of the [...] Read more.
In recent years, the vegetation cover in urban agglomerations has been changing very rapidly due to technogenic influence. Satellite images play a huge role in studying the dynamics of forest vegetation. Special programs are used to process satellite images. The purpose of the study is to analyze forest vegetation within the territory of the Tomsk agglomeration based on Landsat remote sensing data for the period from 1990 to 2022. The novelty of the study is explained by the development of a unique program code for the analysis of Landsat satellite data on the previously unexplored territory of the Tomsk agglomeration with the prospect of moving to the scale of the entire state in the future. In this study, the authors present an algorithm implemented in Python to quantify the change in the area of vegetation in an urban agglomeration using Landsat multispectral data. The tool allows you to read space images, calculate spectral indices (NDVI, UI, NDWI), and perform statistical processing of interpretation results. The created tool was applied to study the dynamics of vegetation within the Tomsk urban agglomeration during the period 1990–2022. Key findings and conclusions: (1) The non-forest areas increased from 1990 to 1999 and from 2013 to 2022. It is very likely that this is due to the deterioration of the standard of living in the country during these periods. The first time interval corresponds to the post-Soviet period and the devastation in the economy in the 1990s. The second period corresponds to the implementation and strengthening of sanctions pressure on the Russian Federation. (2) The area of territories inhabited by people has been steadily falling since 1990. This is due to the destruction of collective agriculture in the Russian Federation and the outflow of the population from the surrounding rural settlements to Tomsk and Seversk. Full article
Show Figures

Figure 1

Figure 1
<p>The study area of Tomsk agglomeration.</p>
Full article ">Figure 2
<p>Flowchart of the research method.</p>
Full article ">Figure 3
<p>NDVI map.</p>
Full article ">Figure 4
<p>UI map.</p>
Full article ">Figure 5
<p>Mapping of forest vegetation in the Tomsk agglomeration. Legend: green for vegetation, purple for urban areas, black for other areas: (<b>a</b>)—1990 year; (<b>b</b>)—1999 year; (<b>c</b>)—2007 year; (<b>d</b>)—2013 year; (<b>e</b>)—2020 year; (<b>f</b>)—2022 year. Red box is maximum changes after initial state.</p>
Full article ">Figure 6
<p>Scatter plots of land cover types in the Tomsk agglomeration in the spectral space of NDBI and UI.</p>
Full article ">
13 pages, 3077 KiB  
Communication
Inversion of Wind and Temperature from Low SNR FPI Interferograms
by Yafei Wei, Sheng-Yang Gu, Zhenlin Yang, Cong Huang, Na Li, Guoyuan Hu and Xiankang Dou
Remote Sens. 2023, 15(7), 1934; https://doi.org/10.3390/rs15071934 - 4 Apr 2023
Cited by 1 | Viewed by 1301
Abstract
The temperature and wind in the middle and upper atmosphere can be obtained by recording the Doppler shift and broadening of the airglow emission, which is reflected by the interference ring from a ground-based Fabry–Perot interferometer (FPI) system. FPI observations are highly susceptible [...] Read more.
The temperature and wind in the middle and upper atmosphere can be obtained by recording the Doppler shift and broadening of the airglow emission, which is reflected by the interference ring from a ground-based Fabry–Perot interferometer (FPI) system. FPI observations are highly susceptible to weather and the external environment, which seriously affect the signal-to-noise ratio (SNR) of FPI interferograms. An SNR can significantly increase errors in determining the center of the interferogram, leading to inaccurate wind and temperature inversions. The calculation shows that the wind inversion from the interferogram decreases and the temperature increases for larger central errors. In this paper, we propose the maximum standard deviation method (MSDM) with high accuracy and robustness to determine the interference ring center. The performance of the MSDM is better achieved by using more than 100 1D interferogram bins to determine the center of interferograms. The robustness of the MSDM is investigated by computing numerous simulated interferograms with white Gaussian noise and Poisson noise, and compared with the two algorithms of binarization and peak fitting, which are usually used to invert wind and temperature from the interference ring of FPI. The results show that MSDM has higher accuracy and robustness than the other two algorithms. We also simulate the distortion interferogram when the FPI may be illuminated by inhomogeneous background light, which can introduce additional errors in wind and temperature, and the MSDM still performs better. Finally, we invert the wind and temperature from the real airglow interferogram by the Kelan (38.7°N, 111.6°E) FPI, which shows that both the wind and temperature inverted by MSDM better agree well with the FPI product than the other two algorithms. Therefore, the MSDM helps to improve the accuracy and stability to invert the wind and temperature. Full article
Show Figures

Figure 1

Figure 1
<p>Simulated interferogram with V = 50 m/s, T = 600 K.</p>
Full article ">Figure 2
<p>(<b>a</b>) Binarization and (<b>b</b>) peak fitting.</p>
Full article ">Figure 3
<p>(<b>a</b>) Wind and (<b>b</b>) temperature versus the deviation between the real center and the center used.</p>
Full article ">Figure 4
<p>The analysis process of the MSDM.</p>
Full article ">Figure 5
<p>The deviation of the interferogram center using the MSDM with different numbers of bins.</p>
Full article ">Figure 6
<p>The top, middle, and bottom rows show the center deviation, wind, and temperature, respectively. The left, middle, and right columns are calculated using the MSDM, binarization, and peak fitting, respectively. The wind and temperature for each method use the center with deviations in the first column.</p>
Full article ">Figure 7
<p>Same as <a href="#remotesensing-15-01934-f006" class="html-fig">Figure 6</a>, but for Poisson noise.</p>
Full article ">Figure 8
<p>(<b>a</b>) Distorted FPI airglow interferogram (Kelan, Day 312 of 2011) and (<b>b</b>) simulated distortion interferogram (<span class="html-italic">K</span> = 0.1).</p>
Full article ">Figure 9
<p>The center deviation using the three algorithms.</p>
Full article ">Figure 10
<p>Inversion with airglow interferograms on day 17 of 2012. Wind and temperature errors and signal intensity provided by FPI products manufactured by NCAR.</p>
Full article ">
18 pages, 8822 KiB  
Article
Data Comparison and Cross-Calibration between Level 1 Products of DPC and POSP Onboard the Chinese GaoFen-5(02) Satellite
by Xuefeng Lei, Zhenhai Liu, Fei Tao, Hao Dong, Weizhen Hou, Guangfeng Xiang, Lili Qie, Binghuan Meng, Congfei Li, Feinan Chen, Yanqing Xie, Miaomiao Zhang, Lanlan Fan, Liangxiao Cheng and Jin Hong
Remote Sens. 2023, 15(7), 1933; https://doi.org/10.3390/rs15071933 - 4 Apr 2023
Cited by 2 | Viewed by 2158
Abstract
The Polarization CrossFire (PCF) suite onboard the Chinese GaoFen-5(02) satellite has been sophisticatedly composed by the Particulate Observing Scanning Polarimeter (POSP) and the Directional Polarimetric Camera (DPC). Among them, DPC is a multi-angle sequential measurement polarization imager, while POSP is a cross-track scanning [...] Read more.
The Polarization CrossFire (PCF) suite onboard the Chinese GaoFen-5(02) satellite has been sophisticatedly composed by the Particulate Observing Scanning Polarimeter (POSP) and the Directional Polarimetric Camera (DPC). Among them, DPC is a multi-angle sequential measurement polarization imager, while POSP is a cross-track scanning simultaneous polarimeter with corresponding radiometric and polarimetric calibrators, which can theoretically be used for cross comparison and calibration with DPC. After the data preprocessing of these two sensors, we first select local homogeneous cluster scenes by calculating the local variance-to-mean ratio in DPC’s Level 1 product projection grids to reduce the influence of scale differences and geometry misalignment between DPC and POSP. Then, taking the observation results after POSP data quality assurance as the abscissa and taking the DPC observation results under the same wavelength band and geometric conditions as the same ordinate, a two-dimensional radiation/polarization feature space is established. Results show that the normalized top of the atmosphere (TOA) radiances of DPC and POSP processed data at the nadir are linearly correlated. The normalized TOA radiance root mean square errors (RMSEs) look reasonable in all common bands. The DPC and POSP normalized radiance ratios in different viewing zenith angle ranges at different times reveal the temporal drift of the DPC relative radiation response. The RMSEs, mean absolute errors (MAEs), relative errors (REs), and scatter percentage of DPC degree of linear polarization (DoLP) falling within the expected error (EE = ±0.02) of POSP measured DoLP are better than 0.012, 0.009, 0.066, and 91%, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>The sensor assembly of the PCF suite installed on the GaoFen-5(02) satellite.</p>
Full article ">Figure 2
<p>The layout diagram of the POSP onboard calibrators.</p>
Full article ">Figure 3
<p>The sampling schematic diagram of PCF [<a href="#B27-remotesensing-15-01933" class="html-bibr">27</a>].</p>
Full article ">Figure 4
<p>Normalized relative spectral responses of DPC and POSP common bands.</p>
Full article ">Figure 5
<p>The linear fitting coefficients of DPC and POSP normalized radiance in common spectral bands at nadir. The data pairs are a collection selected from May 2022 of (<b>a</b>) 443 nm; (<b>b</b>) 490 nm; (<b>c</b>) 670 nm; and (<b>d</b>) 865 nm bands in DPC and POSP Level 1 products. Yellow solid and red dashed lines are the 1:1 lines and fit lines, respectively. The reason for the fewer matching data points is that the amount of raw data points at nadir is small.</p>
Full article ">Figure 5 Cont.
<p>The linear fitting coefficients of DPC and POSP normalized radiance in common spectral bands at nadir. The data pairs are a collection selected from May 2022 of (<b>a</b>) 443 nm; (<b>b</b>) 490 nm; (<b>c</b>) 670 nm; and (<b>d</b>) 865 nm bands in DPC and POSP Level 1 products. Yellow solid and red dashed lines are the 1:1 lines and fit lines, respectively. The reason for the fewer matching data points is that the amount of raw data points at nadir is small.</p>
Full article ">Figure 6
<p>The time-varying characteristics of DPC and POSP normalized radiance linear fitting results (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>A</mi> </mrow> <mrow> <mi>k</mi> </mrow> </msup> </mrow> </semantics></math>; (<b>b</b>) R<sup>2</sup>; and (<b>c</b>) RMSE in common spectral bands at nadir. These data pairs are selected from collections from November 2021 to May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 7
<p>The linear fitting slope and corresponding quadratic fitting curve of DPC and POSP normalized radiance in common spectral bands with different post-launch days and different VZA ranges. The data pairs are collected in each day selected from October 2021 to July 2022 of (<b>a</b>) 443 nm; (<b>b</b>) 490 nm; (<b>c</b>) 670 nm; and (<b>d</b>) 865 nm bands in DPC and POSP Level 1 products.</p>
Full article ">Figure 8
<p>The ratio of DPC and POSP normalized radiance in the 443 nm band with different VZAs. These data pairs are selected for collection from (<b>a</b>) November 2021; (<b>b</b>) December 2021; (<b>c</b>) January 2022; (<b>d</b>) February 2022; (<b>e</b>) April 2022; and (<b>f</b>) May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 9
<p>The ratio of DPC to POSP normalized radiance in the 865 nm band with different VZAs. These data pairs are selected for collection from (<b>a</b>) November 2021; (<b>b</b>) December 2021; (<b>c</b>) January 2022; (<b>d</b>) February 2022; (<b>e</b>) April 2022; and (<b>f</b>) May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 10
<p>The ratio of DPC to POSP normalized radiance in the 490 nm band with different VZAs. These data pairs are selected for collection from (<b>a</b>) November 2021; (<b>b</b>) December 2021; (<b>c</b>) January 2022; (<b>d</b>) February 2022; (<b>e</b>) April 2022; and (<b>f</b>) May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 11
<p>The ratio of DPC to POSP normalized radiance in the 670 nm band with different VZAs. These data pairs are selected for collection from (<b>a</b>) November 2021; (<b>b</b>) December 2021; (<b>c</b>) January 2022; (<b>d</b>) February 2022; (<b>e</b>) April 2022; and (<b>f</b>) May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 11 Cont.
<p>The ratio of DPC to POSP normalized radiance in the 670 nm band with different VZAs. These data pairs are selected for collection from (<b>a</b>) November 2021; (<b>b</b>) December 2021; (<b>c</b>) January 2022; (<b>d</b>) February 2022; (<b>e</b>) April 2022; and (<b>f</b>) May 2022 in DPC and POSP Level 1 products.</p>
Full article ">Figure 12
<p>The linear fitting and statistical distribution results of DPC and POSP DoLP in common polarimetric bands. The data pairs are selected for collection from October 2021 to May 2022 for (<b>a</b>) and (<b>d</b>) 490 nm; (<b>b</b>) and (<b>e</b>) 670 nm; and (<b>c</b>,<b>f</b>) 865 nm bands in DPC and POSP Level 1 products. Yellow solid, green, and red dashed lines are the 1:1 lines, EE envelope lines, and fit lines, respectively.</p>
Full article ">Figure 12 Cont.
<p>The linear fitting and statistical distribution results of DPC and POSP DoLP in common polarimetric bands. The data pairs are selected for collection from October 2021 to May 2022 for (<b>a</b>) and (<b>d</b>) 490 nm; (<b>b</b>) and (<b>e</b>) 670 nm; and (<b>c</b>,<b>f</b>) 865 nm bands in DPC and POSP Level 1 products. Yellow solid, green, and red dashed lines are the 1:1 lines, EE envelope lines, and fit lines, respectively.</p>
Full article ">
19 pages, 2390 KiB  
Technical Note
Radiometric Terrain Flattening of Geocoded Stacks of SAR Imagery
by Piyush S. Agram, Michael S. Warren, Scott A. Arko and Matthew T. Calef
Remote Sens. 2023, 15(7), 1932; https://doi.org/10.3390/rs15071932 - 4 Apr 2023
Cited by 3 | Viewed by 2467
Abstract
We have described an efficient approach to radiometrically flatten geocoded stacks of calibrated synthetic aperture radar (SAR) data for terrain-related effects. We have used simulation to demonstrate that, for the Sentinel-1 mission, one static radiometric terrain-flattening factor derived from actual SAR imaging metadata [...] Read more.
We have described an efficient approach to radiometrically flatten geocoded stacks of calibrated synthetic aperture radar (SAR) data for terrain-related effects. We have used simulation to demonstrate that, for the Sentinel-1 mission, one static radiometric terrain-flattening factor derived from actual SAR imaging metadata per imaging geometry is sufficient for flattening interferometrically compliant stacks of SAR data. We have quantified the loss of precision due to the application of static flattening factors, and show that these are well below the stated requirements of change-detection algorithms. Finally, we have discussed the implications of applying radiometric terrain flattening to geocoded SAR data instead of the traditional approach of flattening data provided in the original SAR image geometry. The proposed approach allows for efficient and consistent generation of five different Committee of Earth-Observation Satellites (CEOS) Analysis-Ready Dataset (ARD) families—Geocoded Single-Look Complex (GSLC), Interferometric Radar (InSAR), Normalized Radar Backscatter (NRB), Polarimetric Radar (POL) and Ocean Radar Backscatter (ORB) from SAR missions in a common framework. Full article
(This article belongs to the Section Engineering Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Normalization area relations for different SAR calibration levels of a single triangular DEM facet. The altitude of the DEM points <math display="inline"><semantics> <msub> <mi>T</mi> <mn>00</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>T</mi> <mn>01</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>T</mi> <mn>10</mn> </msub> </semantics></math> have been exaggerated for clarity. The normal vectors to the reference ellipsoid and the DEM facet are shown as dashed lines. The slant-range plane passes through <math display="inline"><semantics> <msub> <mi>T</mi> <mi>c</mi> </msub> </semantics></math>, but we have projected the plane upwards for clarity. <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>c</mi> <mi>c</mi> </mrow> </msub> </semantics></math> is a point on the reference ellipsoid that maps to the same slant-range coordinates as <math display="inline"><semantics> <msub> <mi>T</mi> <mi>c</mi> </msub> </semantics></math>, and the constant slant-range curve connecting these two points is shown. This image is comparable to Figures 2 and 5 of [<a href="#B4-remotesensing-15-01932" class="html-bibr">4</a>]. The unit vectors <math display="inline"><semantics> <mover accent="true"> <mi>l</mi> <mo>^</mo> </mover> </semantics></math>, <math display="inline"><semantics> <mover accent="true"> <mi>v</mi> <mo>^</mo> </mover> </semantics></math> and <math display="inline"><semantics> <mover accent="true"> <mi>c</mi> <mo>^</mo> </mover> </semantics></math> are used for defining interferometric baselines and are described in greater detail in <a href="#app1-remotesensing-15-01932" class="html-app">Appendix A</a>.</p>
Full article ">Figure 2
<p>Footprints of the three stacks of SAR metadata that were used in this manuscript overlaid on the 1 km GLOBE DEM—Sentinel-1 burst IW1-0151226 over Big Bear in California, Sentinel-1 burst IW3-015131 over the ocean and ALOS Track 216, Frame 740 over Long Valley in California.</p>
Full article ">Figure 3
<p>(<b>Left</b>) Peak-to-peak variation in <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>σ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> in dB, (<b>Middle</b>) Mean projection angle (<math display="inline"><semantics> <msub> <mi>ψ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>j</mi> </mrow> </msub> </semantics></math>) and (<b>Right</b>) Mean local incidence angle (<math display="inline"><semantics> <msub> <mi>θ</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> </semantics></math>) as a function of <math display="inline"><semantics> <msub> <mi>h</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>h</mi> <mn>2</mn> </msub> </semantics></math> in meters of a single facet corresponding to a stack of 58 Sentinel-1 acquisitions corresponding to burst footprint IW1-0151226. The bright line in (<b>Left</b>) corresponds to <math display="inline"><semantics> <mrow> <msub> <mi>ψ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>j</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. Plots have been masked for the region corresponding to <math display="inline"><semantics> <mrow> <mo form="prefix">cos</mo> <msub> <mi>θ</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> <mo>&gt;</mo> <msup> <mo form="prefix">cos</mo> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mfenced separators="" open="(" close=")"> <mn>0.05</mn> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Scatter plots of (<b>Left</b>) along-track baseline (<math display="inline"><semantics> <msubsup> <mi>B</mi> <mi>v</mi> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math>) in meters vs. perpendicular baseline (<math display="inline"><semantics> <msubsup> <mi>B</mi> <mrow> <mo>⊥</mo> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math>) in meters and (<b>Right</b>) parallel baseline (<math display="inline"><semantics> <msubsup> <mi>B</mi> <mrow> <mo>‖</mo> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math>) in meters vs. perpendicular baseline (<math display="inline"><semantics> <msubsup> <mi>B</mi> <mrow> <mo>⊥</mo> </mrow> <mrow> <mi>r</mi> <mi>e</mi> <mi>f</mi> </mrow> </msubsup> </semantics></math>) in meters corresponding to ALOS-1 stack w.r.t the imaging geometry of 2006-06-30 acquisition for <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mn>1</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>(<b>Left</b>) Peak-to-peak variation in <math display="inline"><semantics> <msub> <mi>A</mi> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> </msub> </semantics></math>, (<b>Middle</b>) Mean projection angle (<math display="inline"><semantics> <msub> <mi>ψ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>j</mi> </mrow> </msub> </semantics></math>) and (<b>Right</b>) Mean local incidence angle (<math display="inline"><semantics> <msub> <mi>θ</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> </semantics></math>) as a function of <math display="inline"><semantics> <msub> <mi>h</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>h</mi> <mn>2</mn> </msub> </semantics></math> in meters of a single facet corresponding to a stack of 34 ALOS-1 acquisitions corresponding to Path 216, Frame 740. The bright line in (<b>Left</b>) corresponds to <math display="inline"><semantics> <mrow> <msub> <mi>ψ</mi> <mrow> <mi>p</mi> <mi>r</mi> <mi>j</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>90</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. Plots have been masked for the region corresponding to <math display="inline"><semantics> <mrow> <mo form="prefix">cos</mo> <msub> <mi>θ</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> <mspace width="3.33333pt"/> <mo>&gt;</mo> <mspace width="3.33333pt"/> <msup> <mo form="prefix">cos</mo> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mfenced separators="" open="(" close=")"> <mn>0.05</mn> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Scatter plot of <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>A</mi> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> </msub> </mrow> </semantics></math> in dB vs. perpendicular baseline (<math display="inline"><semantics> <msub> <mi>B</mi> <mo>⊥</mo> </msub> </semantics></math>) corresponding to ALOS stack used in <a href="#sec3dot2-remotesensing-15-01932" class="html-sec">Section 3.2</a> for three different facets. The values of <math display="inline"><semantics> <msub> <mi>h</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>h</mi> <mn>2</mn> </msub> </semantics></math> for the facet, along with the resulting local incidence angle <math display="inline"><semantics> <msub> <mi>θ</mi> <mrow> <mi>i</mi> <mi>n</mi> <mi>c</mi> </mrow> </msub> </semantics></math> are also shown.</p>
Full article ">Figure 7
<p>Histogram of the standard deviation of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>σ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> for all pixels in the open-ocean burst off the California coast with default oversampling factor (<b>Left</b>) and an oversampling factor multiple of 2 (<b>Right</b>). The results from an oversampling factor of 2 match our simulations from <a href="#sec3dot1-remotesensing-15-01932" class="html-sec">Section 3.1</a>.</p>
Full article ">Figure 8
<p>Histogram of the standard deviation of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>σ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> for all pixels in a rugged terrain burst over Big Bear, California with default oversampling factor (<b>Left</b>) and an oversampling factor multiple of 4 (<b>Right</b>). Data were masked with the shadow–layover mask before the histograms were estimated, but the long tail of the histogram for the rugged terrain burst indicates that there could be other subtle processing effects to account for in steep terrain.</p>
Full article ">Figure 9
<p>Standard deviation of the ratio of <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>T</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>σ</mi> <mrow> <mn>0</mn> <mo>,</mo> <mi>E</mi> </mrow> </msub> </mrow> </semantics></math> with masked out regions in white for default oversampling factor (<b>Left</b>) and an oversampling multiple of 4 (<b>Middle</b>). Corresponding DEM over a rugged 15 km × 15 km area near Big Bear, California (<b>Right</b>). An oversampling multiple of 4 reduces the observed processing error but does not eliminate the correlation with topography.</p>
Full article ">
25 pages, 18095 KiB  
Article
Improving Semantic Segmentation of Roof Segments Using Large-Scale Datasets Derived from 3D City Models and High-Resolution Aerial Imagery
by Florian L. Faltermeier, Sebastian Krapf, Bruno Willenborg and Thomas H. Kolbe
Remote Sens. 2023, 15(7), 1931; https://doi.org/10.3390/rs15071931 - 4 Apr 2023
Cited by 5 | Viewed by 3731
Abstract
Advances in deep learning techniques for remote sensing as well as the increased availability of high-resolution data enable the extraction of more detailed information from aerial images. One promising task is the semantic segmentation of roof segments and their orientation. However, the lack [...] Read more.
Advances in deep learning techniques for remote sensing as well as the increased availability of high-resolution data enable the extraction of more detailed information from aerial images. One promising task is the semantic segmentation of roof segments and their orientation. However, the lack of annotated data is a major barrier for deploying respective models on a large scale. Previous research demonstrated the viability of the deep learning approach for the task, but currently, published datasets are small-scale, manually labeled, and rare. Therefore, this paper extends the state of the art by presenting a novel method for the automated generation of large-scale datasets based on semantic 3D city models. Furthermore, we train a model on a dataset 50 times larger than existing datasets and achieve superior performance while applying it to a wider variety of buildings. We evaluate the approach by comparing networks trained on four dataset configurations, including an existing dataset and our novel large-scale dataset. The results show that the network performance measured as intersection over union can be increased from 0.60 for the existing dataset to 0.70 when the large-scale model is applied on the same region. The large-scale model performs superiorly even when applied to more diverse test samples, achieving 0.635. The novel approach contributes to solving the dataset bottleneck and consequently to improving semantic segmentation of roof segments. The resulting remotely sensed information is crucial for applications such as solar potential analysis or urban planning. Full article
Show Figures

Figure 1

Figure 1
<p>A graphical representation of the methodology used in this study. Numbers to the right point to the respective sections in this article where further details can be found.</p>
Full article ">Figure 2
<p>A graphical example of the approach to derive two-dimensional roof segment labels from 3D building models for the case of a simple gable roof. (<b>Left</b>): Exemplary 3D building models from the dataset. Center: The 3D representation of a building with the roof segments’ normal vectors. (<b>Right</b>): The resulting roof segment labels after projection onto the two-dimensional plane and assignation of an orientation class according to a chosen mapping from the continuous azimuth to discrete bins. The axes <span class="html-italic">x</span> and <span class="html-italic">y</span> refer to longitude and latitude, respectively.</p>
Full article ">Figure 3
<p>Data split of the small-scale datasets <span class="html-italic">small-manu</span> and <span class="html-italic">small-auto</span> (identical number, location, and size of training samples) with areas containing training (train), validation (val), and test data. The datasets are based on the German village of Wartenberg. Its location within Bavaria and Germany, respectively, is indicated in the overview maps.</p>
Full article ">Figure 4
<p>Data split of the large-scale dataset <span class="html-italic">large-auto</span> with areas containing training (train), validation (val), and test data. One detailed map for each region (<b>a</b>–<b>e</b>) and one overview map that illustrates their spatial relation: (<b>a</b>) Freising in the very north; (<b>b</b>) Erding in the very east; (<b>c</b>,<b>d</b>) two areas in Munich centrally; (<b>e</b>) sparsely populated rural area in the south-west.</p>
Full article ">Figure 5
<p>Exemplary comparison of manual and automatic samples from the small-scale datasets. IoU is given. (<b>A</b>) Manual labels identify dormer, auto labels do not; on the other hand, auto labels identify segments missing in manual dataset. (<b>B</b>) Auto labels do not cover roof overhangs and sometimes wrongly represent roof geometry, as for the cross-gabled building to the right.</p>
Full article ">Figure 6
<p>Confusion matrices of the models <span class="html-italic">small-auto</span> and <span class="html-italic">large-auto</span> evaluated on their own test sets. Rows are ground truth, columns are predictions. Rows are normalized to total number of predictions, but do not always sum up to one due to rounding to two digits. Last column shows IoU of each class.</p>
Full article ">Figure 7
<p>Test samples from the small-scale datasets <span class="html-italic">small-manu</span> and <span class="html-italic">small-auto</span> at two locations (<b>A</b>,<b>B</b>), and corresponding predictions from all models. For each location, the upper sample is from <span class="html-italic">small-manu</span> (Google image, manual labels) and the lower sample is from <span class="html-italic">small-auto</span> (LDBV image, auto labels). IoU with respect to labels is given for each prediction.</p>
Full article ">Figure 8
<p>Test samples from the large-scale dataset <span class="html-italic">large-auto</span> at two locations (<b>A</b>,<b>B</b>), and corresponding predictions from all models. IoU with respect to labels is given for each prediction.</p>
Full article ">
11 pages, 5709 KiB  
Communication
A New Algorithm for Ill-Posed Problem of GNSS-Based Ionospheric Tomography
by Debao Wen, Kangyou Xie, Yinghao Tang, Dengkui Mei, Xi Chen and Hanqing Chen
Remote Sens. 2023, 15(7), 1930; https://doi.org/10.3390/rs15071930 - 4 Apr 2023
Cited by 4 | Viewed by 1550
Abstract
Ill-posedness of GNSS-based ionospheric tomography affects the stability and the accuracy of the inversion results. Truncated singular value decomposition (TSVD) is a common algorithm of ionospheric tomography reconstruction. However, the TSVD method usually has low inversion accuracy and reconstruction efficiency. To resolve the [...] Read more.
Ill-posedness of GNSS-based ionospheric tomography affects the stability and the accuracy of the inversion results. Truncated singular value decomposition (TSVD) is a common algorithm of ionospheric tomography reconstruction. However, the TSVD method usually has low inversion accuracy and reconstruction efficiency. To resolve the above problem, a truncated mapping singular value decomposition (TMSVD) algorithm is presented to improve the reconstructed accuracy and computational efficiency. To authenticate the effectiveness and the advantages of the TMSVD algorithm, a numerical test scheme is devised. Finally, ionospheric temporal–spatial variations of the selected reconstructed region are studied using the GNSS observations under different geomagnetic conditions. The reconstructed results of TMSVD can accurately reflect semiannual anomalies, diurnal variations, and geomagnetic storm effects. In contrast with the ionosonde data, it is found that the reconstructed profiles of the TMSVD method are more consistent with than those of the IRI 2016. The study suggests that TMSVD is an efficient algorithm for the tomographic reconstruction of ionospheric electron density (IED). Full article
(This article belongs to the Special Issue Ionosphere Monitoring with Remote Sensing II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparisons of the tomographic results of two algorithms with the simulated IED true values. (<b>a</b>) TMSVD; (<b>b</b>) TSVD.</p>
Full article ">Figure 2
<p>Error statistics of the TMSVD and the TSVD. (<b>a</b>) TMSVD; (<b>b</b>) TSVD.</p>
Full article ">Figure 3
<p>Locations of the selected GNSS and ionosonde stations.</p>
Full article ">Figure 4
<p>The horizontal section of the reconstructed IED distributions using TMSVD at the altitude of 150–650 km. (<b>a</b>) 5:00 UT on 4 January 2022; (<b>b</b>) 5:00 UT on 17 March 2022; (<b>c</b>) 5:00 UT on 17 June 2022; (<b>d</b>) 5:00 UT on 18 September 2022.</p>
Full article ">Figure 5
<p>The three-dimensional IED distributions reconstructed by TMSVD at 05:00 UT. (<b>a</b>) 4 January 2022; (<b>b</b>) 17 March 2022; (<b>c</b>) 17 June 2022; (<b>d</b>) 18 September 2022.</p>
Full article ">Figure 6
<p>The IED diurnal variation at the altitude of 350 km on 25 August 2018. The unit of IED is 10<sup>11</sup> el/m<sup>3</sup>.</p>
Full article ">Figure 7
<p>The three-dimensional IED distributions reconstructed using TMSVD and the IRI 2016 model at 8:00 on 25–26 August 2018.</p>
Full article ">Figure 8
<p>The comparisons of the IED profiles reconstructed by the TMSVD with those obtained from the IRI 2016 and ionosonde in Shaoyang.</p>
Full article ">
17 pages, 4485 KiB  
Article
Recovery of Water Quality and Detection of Algal Blooms in Lake Villarrica through Landsat Satellite Images and Monitoring Data
by Lien Rodríguez-López, Iongel Duran-Llacer, Lisandra Bravo Alvarez, Andrea Lami and Roberto Urrutia
Remote Sens. 2023, 15(7), 1929; https://doi.org/10.3390/rs15071929 - 3 Apr 2023
Cited by 8 | Viewed by 5345
Abstract
Phytoplankton is considered a strong predictor of the environmental quality of lakes, while Chlorophyll-a is an indicator of primary productivity. In this study, 25 LANDSAT images covering the 2014–2021 period were used to predict Chlorophyll-a in the Villarrica lacustrine system. A Chlorophyll-a recovery [...] Read more.
Phytoplankton is considered a strong predictor of the environmental quality of lakes, while Chlorophyll-a is an indicator of primary productivity. In this study, 25 LANDSAT images covering the 2014–2021 period were used to predict Chlorophyll-a in the Villarrica lacustrine system. A Chlorophyll-a recovery algorithm was calculated using two spectral indices (FAI and SABI). The indices that presented the best statistical indicators were the floating algal index (R2 = 0.87) and surface algal bloom index (R2 = 0.59). A multiparametric linear model for Chlorophyll-a estimation was constructed with the indices. Statistical indicators were used to validate the multiple linear regression model used to predict Chlorophyll-a by means of spectral indices, with the following results: a MBE of −0.136 μ, RMSE of 0.055 μ, and NRMSE of 0.019%. All results revealed the strength of the model. It is necessary to raise awareness among the population that carries out activities around the lake in order for them to take policy actions related to water resources in this Chilean lake. Furthermore, it is important to note that this study is the first to address the detection of algal blooms in this Chilean lake through remote sensing. Full article
(This article belongs to the Special Issue Remote Sensing for Marine Environmental Disaster Response)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Behavior of limnological parameters in Lake Villarrica during summer and spring seasons. For the summer months (December, January, and February) and for the spring months (September, October, and November) were considered.</p>
Full article ">Figure 2
<p>Measured Chl-a and Chl-a estimated using the obtained multiple linear regression model.</p>
Full article ">Figure 3
<p>Spatial distribution of FAI in Lake Villarrica.</p>
Full article ">Figure 4
<p>Spatial distribution of SABI in Lake Villarrica.</p>
Full article ">Figure 5
<p>Spatial pattern of Chlorophyll-a arising from multivariate regression resulting from spectral indices through as seen Landsat imagery.</p>
Full article ">Figure 6
<p>(<b>a</b>) Chile in South America; (<b>b</b>) Araucanía Region with topographic profiles, with the study area, including Lake Villarrica, outlined in black; and (<b>c</b>) the land uses of the Villarrica basin.</p>
Full article ">
23 pages, 28261 KiB  
Article
Spatio-Temporal Changes of Mangrove-Covered Tidal Flats over 35 Years Using Satellite Remote Sensing Imageries: A Case Study of Beibu Gulf, China
by Ertao Gao and Guoqing Zhou
Remote Sens. 2023, 15(7), 1928; https://doi.org/10.3390/rs15071928 - 3 Apr 2023
Cited by 3 | Viewed by 2848
Abstract
Tidal flats provide ecosystem services to billions of people worldwide; however, their changing status is largely unknown. Several challenges in the fine extraction of tidal flats using remote sensing techniques, including tide-level and water-edge line changes, exist at present, especially regarding the spatial [...] Read more.
Tidal flats provide ecosystem services to billions of people worldwide; however, their changing status is largely unknown. Several challenges in the fine extraction of tidal flats using remote sensing techniques, including tide-level and water-edge line changes, exist at present, especially regarding the spatial and temporal distribution of mangroves. This study proposed a tidal flats extraction method using a combination of threshold segmentation and tidal-level correction, considering the influence of mangrove changes. We extracted the spatial distribution of tidal flats in Beibu Gulf, Southwest China, from 1987 to 2021 using time-series Landsat and Sentinel-2 images, and further analyzed the dynamic variation characteristics of the total tidal flats, each coastal segment, and the range of erosion and silting. To quantitatively investigate the interaction between tidal flats and mangroves, this study established a regression model based on multi-temporal tidal flats and mangrove data. The results indicated that the overall accuracy of the tidal flat extraction results was 93.9%, and the kappa coefficient was 0.82. The total area of tidal flats in Beibu Gulf decreased by 130 km2 from 1987 to 2021, with an average annual change of −3.7 km2/a. In addition, a negative correlation between the tidal flat change area and mangrove change area in Shankou, Maowei Sea, and Pearl Bay was observed, with correlation coefficients of −0.28, −0.30 and −0.64, respectively. These results demonstrate that the distribution of tidal flats provides a good environment and expansion space for the rapid growth of mangroves. These results can provide references for tidal flats’ resource conservation, ecological health assessment, and vegetation changes in coastal wetlands in China and other countries in Southeast Asia. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area. (<b>a</b>) Pearl Bay Mangrove Reserve. (<b>b</b>) Maowei Sea Mangrove Reserve. (<b>c</b>) Shankou Mangrove Reserve.</p>
Full article ">Figure 2
<p>Comparison of calculation results of <span class="html-italic">NDWI</span> and <span class="html-italic">MNDWI</span>. (<b>a</b>) <span class="html-italic">NDWI</span> of water bodies; (<b>b</b>) <span class="html-italic">MNDWI</span> of water bodies.</p>
Full article ">Figure 3
<p>Instantaneous water-edge line extraction process. (<b>a</b>) Water edge image to be extracted. (<b>b</b>) <span class="html-italic">MNDWI</span> Index. (<b>c</b>) Threshold segmentation binarization. (<b>d</b>) Instantaneous water edge extraction.</p>
Full article ">Figure 4
<p>Schematic diagram of tidal-level correction principle.</p>
Full article ">Figure 5
<p>Comparison of before and after correction of tide level in 1987. (<b>a</b>) Before and after correction of tide level at low tide. (<b>b</b>) Before and after correction of tide level at high tide.</p>
Full article ">Figure 6
<p>Photo of tidal flats and mangroves in the study area.</p>
Full article ">Figure 7
<p>Comparison of multispectral imagery (MSI) and <span class="html-italic">IMFI</span> imagery. (<b>a</b>) Mangroves on MSI. (<b>b</b>) Mangroves on IMFI imagery.</p>
Full article ">Figure 8
<p>Research technical route of this study.</p>
Full article ">Figure 9
<p>Comparison of tidal flats extractions with UQD data in Beihai Coast. (<b>a</b>) The 1987 tidal flats extracted by the method in this paper. (<b>b</b>) The 1987 tidal flats of UQD. (<b>c</b>) The 1995 tidal flats extracted by the method in this paper. (<b>d</b>) The 1995 tidal flats of UQD.</p>
Full article ">Figure 10
<p>Distribution of tidal flats in Beibu Gulf in 1987 and 2021. (<b>a</b>) Distribution of tidal flats in Beibu Gulf in 1987. (<b>b</b>) Distribution of tidal flats in Beibu Gulf in 2021.</p>
Full article ">Figure 11
<p>Statistics of total tidal flat area in Beibu Gulf, 1987–2021.</p>
Full article ">Figure 12
<p>Distribution of spatio-temporal siltation and erosion of tidal flats in the Beihai shore section. (<b>a</b>) Erosion distribution of tidal flats in Beihai shore section from 2008 to 2018. (<b>b</b>) Siltation distribution of tidal flats in Beihai shore section from 2006 to 2008. (<b>c</b>) Siltation distribution of tidal flats in Beihai shore section from 2003 to 2005. (<b>d</b>) Siltation distribution of tidal flats in Beihai shore section from 1990 to 1994.</p>
Full article ">Figure 13
<p>Spatio-temporal siltation erosion distribution of tidal flats in Qinzhou shore section. (<b>a</b>) Erosion distribution of tidal flats in Qinzhou shore section from 2008 to 2021. (<b>b</b>) Erosion distribution of tidal flats in Qinzhou shore section from 1997 to 1998. (<b>c</b>) Siltation distribution of tidal flats in Qinzhou shore section from 1993 to 1997. (<b>d</b>) Erosion distribution of tidal flats in Qinzhou shore section from 1987 to 1993.</p>
Full article ">Figure 14
<p>Spatio-temporal siltation erosion distribution of tidal flats in Fangchenggang shore section. (<b>a</b>) Erosion distribution of tidal flats in Fangchenggang shore section from 2015 to 2017. (<b>b</b>) Siltation distribution of tidal flats in Fangchenggang shore section from 2008 to 2015. (<b>c</b>) Siltation distribution of tidal flats in Fangchenggang shore section from 1998 to 2001. (<b>d</b>) Erosion distribution of tidal flats in Fangchenggang shore section from 1987 to 1993.</p>
Full article ">Figure 15
<p>Spatio-temporal distribution of mangroves in Maowei Sea. (<b>a</b>) Distribution of mangroves in Maowei Sea in 1996. (<b>b</b>) Distribution of mangroves in Maowei Sea in 2000. (<b>c</b>) Distribution of mangroves in Maowei Sea in 2005. (<b>d</b>) Distribution of mangroves in Maowei Sea in 2010. (<b>e</b>) Distribution of mangroves in Maowei Sea in 2015. (<b>f</b>) Distribution of mangroves in Maowei Sea in 2021.</p>
Full article ">Figure 16
<p>Latest (2021) tidal flats and mangrove distribution in the study area. (<b>a</b>) Pearl Bay tidal flats and mangrove distribution. (<b>b</b>) Maowei Sea Mangrove Reserve tidal flats and mangrove distribution. (<b>c</b>) Shankou Mangrove Reserve tidal flats and mangrove distribution.</p>
Full article ">Figure 17
<p>Map of land stock and annual growth rate in Shankou region from 1996 to 2021.</p>
Full article ">Figure 18
<p>Tidal sequence change and centroid transfer of mangrove tidal flat in Shankou area. (<b>a</b>) Interclass changes of mangroves and tidal flats in Shankou area from 1996 to 2000. (<b>b</b>) Interclass changes of mangroves and tidal flats in Shankou area from 2000 to 2005. (<b>c</b>) Interclass changes of mangroves and tidal flats in Shankou area from 2005 to 2010. (<b>d</b>) Interclass changes of mangroves and tidal flats in Shankou area from 2010 to 2015. (<b>e</b>) Interclass changes of mangroves and tidal flats in Shankou area from 2015 to 2021.</p>
Full article ">Figure 19
<p>Map of land stock and annual growth rate in Maowei sea region from 1996 to 2021.</p>
Full article ">Figure 20
<p>Tidal sequence change and centroid transfer of mangrove tidal flat in Maowei sea area. (<b>a</b>) Interclass changes of mangroves and tidal flats in Maowei sea from 1996 to 2000. (<b>b</b>) Interclass changes of mangroves and tidal flats in Maowei sea from 2000 to 2005. (<b>c</b>) Interclass changes of mangroves and tidal flats in Maowei sea from 2005 to 2010. (<b>d</b>) Interclass changes of mangroves and tidal flats in Maowei sea from 2010 to 2015. (<b>e</b>) Interclass changes of mangroves and tidal flats in Maowei sea from 2015 to 2021.</p>
Full article ">Figure 21
<p>Map of land stock and annual growth rate in Pearl bay region from 1996 to 2021.</p>
Full article ">Figure 22
<p>Tidal sequence change and centroid transfer of mangrove tidal flat in Pearl bay. (<b>a</b>) Interclass changes of mangroves and tidal flats in Pearl bay from 1996 to 2000. (<b>b</b>) Interclass changes of mangroves and tidal flats in Pearl bay from 2000 to 2005. (<b>c</b>) Interclass changes of mangroves and tidal flats in Pearl bay from 2005 to 2010. (<b>d</b>) Interclass changes of mangroves and tidal flats in Pearl bay from 2010 to 2015. (<b>e</b>) Interclass changes of mangroves and tidal flats in Pearl bay from 2015 to 2021.</p>
Full article ">
23 pages, 30351 KiB  
Article
Comparing the Capability of Sentinel-2 and Landsat 9 Imagery for Mapping Water and Sandbars in the River Bed of the Lower Tagus River (Portugal)
by Romeu Gerardo and Isabel P. de Lima
Remote Sens. 2023, 15(7), 1927; https://doi.org/10.3390/rs15071927 - 3 Apr 2023
Cited by 5 | Viewed by 3695
Abstract
Mapping river beds to identify water and sandbars is a crucial task for understanding the morphology and hydrodynamics of rivers and their ecological conditions. The main difficulties of this task so far have been the limitations of conventional approaches, which are generally costly [...] Read more.
Mapping river beds to identify water and sandbars is a crucial task for understanding the morphology and hydrodynamics of rivers and their ecological conditions. The main difficulties of this task so far have been the limitations of conventional approaches, which are generally costly (e.g., equipment, time- and human resource-demanding) and have poor flexibility to deal with all river conditions. Currently, alternative approaches rely on remote sensing techniques, which offer innovative tools for mapping water bodies in a quick and cost-effective manner based on relevant spectral indices. This study aimed to compare the capability of using imagery from the Sentinel-2 and newly launched Landsat 9 satellite to achieve this goal. For a segment of the Lower Tagus River (Portugal) with conditions of very low river discharge, comparison of the Normalized Difference Water Index, Modified Normalized Difference Water Index, Augmented Normalized Difference Water Index, and Automated Water Extraction Index calculated from the imagery of the two satellites shows that the two satellites’ datasets and mapping were consistent and therefore could be used complementarily. However, the results highlighted the need to classify satellite imagery based on index-specific classification decision values, which is an important factor in the quality of the information extracted. Full article
(This article belongs to the Special Issue Initial Understanding of Landsat-9 Capabilities and Applications)
Show Figures

Figure 1

Figure 1
<p>Identification of the river segment investigated in this study (<b>left</b>) and its location in the Tagus River (<b>right</b>). The view of the river segment, which captured a condition of very low river discharge that occurred in the summer of 2022, reveals the extensive presence of sandbars along the river bed.</p>
Full article ">Figure 2
<p>Data from Almourol hydrometric station (17G/02H), which was located just upstream of the studied river segment: (<b>a</b>) annual discharge time series for the period 1990/91–2019/2020; (<b>b</b>) daily discharge time series, for the period 1 October 2021–25 October 2022. A period of exceptional low river discharge occurred in the summer of 2022.</p>
Full article ">Figure 3
<p>Flowchart illustrating the methodological steps that were carried out in this study for mapping water and non-water (sandbars) surfaces using Sentinel-2 and Landsat 9 satellite imagery.</p>
Full article ">Figure 4
<p>Using Sentinel-2 imagery. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) Spectral water indices maps of a segment of the Tagus River (Portugal) for NDWI, MNDWI, ANDWI, AWEIsh, and AWEInsh; the darker colors indicate non-water features. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) Corresponding data histograms. The data spatial resolution is 20 m. The date of observation was 8 July 2022.</p>
Full article ">Figure 4 Cont.
<p>Using Sentinel-2 imagery. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) Spectral water indices maps of a segment of the Tagus River (Portugal) for NDWI, MNDWI, ANDWI, AWEIsh, and AWEInsh; the darker colors indicate non-water features. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) Corresponding data histograms. The data spatial resolution is 20 m. The date of observation was 8 July 2022.</p>
Full article ">Figure 5
<p>Using Landsat 9 imagery. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>) Spectral water indices maps of a segment of the Tagus River (Portugal) for NDWI, MNDWI, ANDWI, AWEIsh, and AWEInsh; the darker colors indicate non-water features. (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>) Corresponding data histograms. The data spatial resolution is 30 m. The date of observation was 7 July 2022.</p>
Full article ">Figure 6
<p>Correlation analysis of pixel=based water index values calculated from Sentinel-2 and Landsat 9 imagery: (<b>a</b>) NDWI; (<b>b</b>) MNDWI; (<b>c</b>) ANDWI; (<b>d</b>) AWEIsh; and (<b>e</b>) AWEInsh. The corresponding linear regression lines, equations, and correlation coefficients are shown.</p>
Full article ">Figure 7
<p>Comparison of classification results obtained for the NDWI, MNDWI, ANDWI, AWEIsh, and AWEInsh calculated from Sentinel-2 (S2) and Landsat 9 (L9) data, applying manual decision and the minimum within-class variance (Otsu method) criterion: (<b>a</b>) J–M distance found for the class segmentation decision values; (<b>b</b>) water surface percentage area in the studied segment of the Lower Tagus River (Portugal). The classes are for water and non-water (i.e., sandbars) surfaces. The 1:1 line is shown for reference. The remote sensing data spatial resolution is 20 m for S2 (8 July 2022) and 30 m for L9 (7 July 2022), for similar discharge conditions.</p>
Full article ">Figure 8
<p>Relationship between the percentage area estimated for the water surface in the selected segment of the Lower Tagus River (Portugal) and the water/non-water class segmentation decision values (estimated manually and by the Otsu method) used to classify the spectral water indices: (<b>a</b>) NDWI, MNDWI, and ANDWI; and (<b>b</b>) AWEIsh and AWEInsh. Data are from Sentinel-2 (S2) and Landsat 9 (L9). The remote sensing data spatial resolution is 20 m for S2 (8 July 2022) and 30 m for L9 (7 July 2022), for similar discharge conditions.</p>
Full article ">Figure 9
<p>(<b>a</b>) View of the selected segment of the Lower Tagus River (Portugal) and adjacent areas. (<b>b</b>–<b>f</b>) Water and sandbar (non-water) mapping of the river segment based on the classification of water indices calculated using Sentinel 2 data (8 July 2022). The class separation threshold was estimated by manual decision. Data resolution is 20 m.</p>
Full article ">Figure 10
<p>(<b>a</b>) View of the selected segment of the Lower Tagus River (Portugal) and adjacent areas. (<b>b</b>–<b>f</b>) Water and sandbar (non-water) mapping for the river segment based on the classification of water indices calculated using Landsat 9 data (7 July 2022). The class separation threshold was estimated by manual decision. Data resolution is 30 m.</p>
Full article ">Figure 11
<p>Detailed view of the river bed mapping in a section of the river dominated by the presence of a large single sandbar. The two panels more to left show 20 m Sentinel-2 RGB image (<b>top row</b>) and 30 m Landsat 9 RGB image (<b>bottom row</b>). The other panels show the corresponding mapping that resulted from the analysis of the five water indices (NDWI, MMNDWI, ANDWI, AWEInsh, AWEIsh) using the same criteria applied in <a href="#remotesensing-15-01927-f009" class="html-fig">Figure 9</a> and <a href="#remotesensing-15-01927-f010" class="html-fig">Figure 10</a>.</p>
Full article ">Figure 12
<p>Detailed view of the river bed mapping in a section of the river that exhibits the presence of multiple sandbars. The two panels more to left show 20 m Sentinel-2 RGB image (<b>top row</b>) and 30 m Landsat 9 RGB image (<b>bottom row</b>). The other panels show the corresponding mapping that resulted from the analysis of the five water indices (NDWI, MMNDWI, ANDWI, AWEInsh, AWEIsh) using the same criteria applied in <a href="#remotesensing-15-01927-f009" class="html-fig">Figure 9</a> and <a href="#remotesensing-15-01927-f010" class="html-fig">Figure 10</a>.</p>
Full article ">
20 pages, 9703 KiB  
Article
Temporal and Spatial Change in Vegetation and Its Interaction with Climate Change in Argentina from 1982 to 2015
by Qi Long, Fei Wang, Wenyan Ge, Feng Jiao, Jianqiao Han, Hao Chen, Fidel Alejandro Roig, Elena María Abraham, Mengxia Xie and Lu Cai
Remote Sens. 2023, 15(7), 1926; https://doi.org/10.3390/rs15071926 - 3 Apr 2023
Cited by 5 | Viewed by 4054
Abstract
Studying vegetation change and its interaction with climate change is essential for regional ecological protection. Previous studies have demonstrated the impact of climate change on regional vegetation in South America; however, studies addressing the fragile ecological environment in Argentina are limited. Therefore, we [...] Read more.
Studying vegetation change and its interaction with climate change is essential for regional ecological protection. Previous studies have demonstrated the impact of climate change on regional vegetation in South America; however, studies addressing the fragile ecological environment in Argentina are limited. Therefore, we assessed the vegetation dynamics and their climatic feedback in five administrative regions of Argentina, using correlation analysis and multiple regression analysis methods. The Normalized Difference Vegetation Index 3rd generation (NDVI3g) from Global Inventory Monitoring and Modeling Studies (GIMMS) and climatic data from the Famine Early Warning Systems Network (FEWS NET) Land Data Assimilation System (FLDAS) were processed. The NDVI of the 1982–2015 period in Argentina showed a downward trend, varying from −1.75 to 0.69/decade. The NDVI in Northeast Argentina (NEA), Northwest Argentina (NWA), Pampas, and Patagonia significantly decreased. Precipitation was negatively correlated with the NDVI in western Patagonia, whereas temperature and solar radiation were positively correlated with the NDVI. Extreme precipitation and drought were essential causes of vegetation loss in Patagonia. The temperature (73.09%), precipitation (64.02%), and solar radiation (73.27%) in Pampas, Cuyo, NEA, and NWA were positively correlated with the NDVI. However, deforestation and farming and pastoral activities have caused vegetation destruction in Pampas, NEA, and NWA. Environmental protection policies and deforestation regulations should be introduced to protect the ecological environment. The results of this study clarify the reasons for the vegetation change in Argentina and provide a theoretical reference for dealing with climate change. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geographic distribution of the five administrative regions of Argentina. (<b>b</b>) Land cover in Argentina.</p>
Full article ">Figure 2
<p>(<b>a</b>) Annual average temperature, (<b>b</b>) annual average precipitation, and (<b>c</b>) annual average solar radiation in Argentina from 1982 to 2015.</p>
Full article ">Figure 3
<p>(<b>a</b>) Annual NDVI time series in Argentina from 1982 to 2015. (<b>b</b>) Monthly NDVI time series in Argentina from 1982 to 2015. (<b>c</b>) Mann–Kendall (M-K) test of the NDVI in Argentina from 1982 to 2015. (<b>d</b>) Seasonal NDVI time series in Argentina from 1982 to 2015.</p>
Full article ">Figure 4
<p>(<b>a</b>) Average NDVI. (<b>b</b>) NDVI trend in Argentina from 1982 to 2015.</p>
Full article ">Figure 5
<p>NDVI time series sample at individual pixels in Argentina from 1982 to 2015. (<b>a</b>–<b>i</b>) are the time series of selected sample pixels from 1982 to 2015.</p>
Full article ">Figure 6
<p>(<b>a</b>) Correlation coefficient and (<b>b</b>) significance test of the NDVI and temperature in Argentina from 1982 to 2015. (<b>c</b>) Correlation coefficient and (<b>d</b>) significance test of the NDVI and precipitation in Argentina from 1982 to 2015. (<b>e</b>) Correlation coefficient and (<b>f</b>) significance test of the NDVI and solar radiation in Argentina from 1982 to 2015.</p>
Full article ">Figure 7
<p>(<b>a</b>) The slope of the predicted NDVI (<span class="html-italic">NDVI<sub>pre</sub></span>) in Argentina from 1982 to 2015. (<b>b</b>) The slope of the residual NDVI (<span class="html-italic">NDVI<sub>res</sub></span>) in Argentina from 1982 to 2015.</p>
Full article ">Figure 8
<p>Driving forces of vegetation change in Argentina from 1982 to 2015.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop