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

Next Issue
Volume 16, April-2
Previous Issue
Volume 16, 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 16, Issue 7 (April-1 2024) – 197 articles

Cover Story (view full-size image): This study introduces PrISM, a method for estimating irrigation amounts using remote sensing soil moisture data. It adapts the Antecedent Precipitation Index model through data assimilation. Tested in Catalonia, Spain, for eight consecutive years and during a severe drought in 2023, PrISM accurately identified areas with water restrictions in 2023. It shows good performance, correlating well with in situ data (0.58 to 0.76) and exhibiting a cumulative weekly RMSE between 7 and 11 mm/week. PrISM's effectiveness varies with irrigation techniques, performing well with sprinkler and flood systems but facing challenges with drip irrigation. This paper fills a gap in remote sensing-based irrigation estimation, focusing on detecting significant reductions in water allocations during droughts. 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:
17 pages, 7005 KiB  
Article
Construction of Aerosol Model and Atmospheric Correction in the Coastal Area of Shandong Peninsula
by Kunyang Shan, Chaofei Ma, Jingning Lv, Dan Zhao and Qingjun Song
Remote Sens. 2024, 16(7), 1309; https://doi.org/10.3390/rs16071309 - 8 Apr 2024
Viewed by 964
Abstract
Applying standard aerosol models for atmospheric correction in nearshore coastal waters introduces significant uncertainties due to their inability to accurately represent aerosol characteristics in these regions. To improve the accuracy of remote sensing reflectance (Rrs) products in the nearshore [...] Read more.
Applying standard aerosol models for atmospheric correction in nearshore coastal waters introduces significant uncertainties due to their inability to accurately represent aerosol characteristics in these regions. To improve the accuracy of remote sensing reflectance (Rrs) products in the nearshore waters of the Shandong Peninsula, this study develops an aerosol model based on aerosol data collected from the Mu Ping site in the coastal area of the Shandong Peninsula, enabling tailored atmospheric correction for this specific region. Given the pronounced seasonal variations in aerosol optical properties, monthly aerosol models were developed. The monthly aerosol model is derived using the average values of aerosol microphysical properties. Compared to the standard aerosol model, this model is more effective in characterizing the absorption and scattering characteristics of aerosols in the study area. Corresponding lookup tables for the aerosol model were created and integrated into the NIR-SWIR atmospheric correction algorithm. According to the accuracy evaluation indexes of RMSD, MAE, and UPD, it can be found that the atmospheric correction results of the aerosol model established in this paper are better than those of the standard aerosol model, especially in the 547 nm band. It demonstrates that the new aerosol model outperforms the standard model in atmospheric correction performance. With the increasing availability of aerosol observational data, the aerosol model is expected to become more accurate and applicable to other satellite missions. Full article
(This article belongs to the Special Issue Aerosol and Atmospheric Correction)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The geographical location of the Mu Ping station.</p>
Full article ">Figure 2
<p>Comparison of aerosol size distributions obtained by the GRASP algorithm and AERONET; (<b>a</b>) Beijing_PKU, (<b>b</b>) Socheongcho.</p>
Full article ">Figure 3
<p>Monthly average particle size distribution.</p>
Full article ">Figure 4
<p>Real and imaginary parts of the monthly average complex refractive index at 442 nm.</p>
Full article ">Figure 5
<p>An example illustrating the least squares fitting relationship between variables <math display="inline"><semantics> <mrow> <mi mathvariant="normal">ln</mi> <mo>⁡</mo> <mo>(</mo> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi mathvariant="normal">ln</mi> <mo>⁡</mo> <mo>(</mo> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>a</mi> <mi>s</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> for four MODIS bands, using the aerosol model for June under specific geometric conditions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>36</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>v</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>66</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <msup> <mrow> <mn>70</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>). The color of each point reflects the aerosol optical depth at 550 nm (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mn>550</mn> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 5 Cont.
<p>An example illustrating the least squares fitting relationship between variables <math display="inline"><semantics> <mrow> <mi mathvariant="normal">ln</mi> <mo>⁡</mo> <mo>(</mo> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi mathvariant="normal">ln</mi> <mo>⁡</mo> <mo>(</mo> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>a</mi> <mi>s</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> for four MODIS bands, using the aerosol model for June under specific geometric conditions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>36</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>v</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>66</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <msup> <mrow> <mn>70</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>). The color of each point reflects the aerosol optical depth at 550 nm (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mn>550</mn> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 6
<p>An example illustrating the least squares fitting relationship between variables <math display="inline"><semantics> <mrow> <mi mathvariant="normal">t</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> for four MODIS bands, using the aerosol model for June under specific geometric conditions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>36</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>).</p>
Full article ">Figure 6 Cont.
<p>An example illustrating the least squares fitting relationship between variables <math display="inline"><semantics> <mrow> <mi mathvariant="normal">t</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> for four MODIS bands, using the aerosol model for June under specific geometric conditions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>36</mn> </mrow> <mrow> <mo>∘</mo> </mrow> </msup> </mrow> </semantics></math>).</p>
Full article ">Figure 7
<p>Atmospheric corrected <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> products (443, 547, 645 nm) of MODIS-Aqua over the nearshore of Shandong Peninsula region on 26 September 2020. (new aerosol model: (<b>a</b>–<b>c</b>), NASA aerosol model: (<b>d</b>–<b>f</b>)).</p>
Full article ">Figure 8
<p>Comparison between satellite-retrieved spectra and in situ measurements at the Mu Ping site.</p>
Full article ">Figure 9
<p>Scatter plots comparing the satellite-derived remote sensing reflectance (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>) using the NIR-SWIR atmospheric correction method with the in situ <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> at three MODIS bands (443 nm (<b>left</b>), 547 nm (<b>middle</b>), and 645 nm (<b>right</b>)). The evaluation metrics for the atmospheric correction results of the nine MODIS bands are specifically presented in <a href="#remotesensing-16-01309-t001" class="html-table">Table 1</a>.</p>
Full article ">
0 pages, 9407 KiB  
Article
How Land Use Transitions Contribute to the Soil Organic Carbon Accumulation from 1990 to 2020
by Zihui Zhang, Lang Xia, Zifei Zhao, Fen Zhao, Guanyu Hou, Shixin Wu, Xiao Sun, Shangrong Wu, Peng Yang and Yan Zha
Remote Sens. 2024, 16(7), 1308; https://doi.org/10.3390/rs16071308 - 8 Apr 2024
Cited by 1 | Viewed by 881
Abstract
Soil organic carbon stock (SOCS) changes caused by land use changes are still unclear, and understanding this response is essential for many environmental policies and land management practices. In this study, we investigated the temporal–spatial and vertical distribution characteristics of the SOCS in [...] Read more.
Soil organic carbon stock (SOCS) changes caused by land use changes are still unclear, and understanding this response is essential for many environmental policies and land management practices. In this study, we investigated the temporal–spatial and vertical distribution characteristics of the SOCS in the Western Sichuan Plateau (WSP) using the sparrow search algorithm–random forest regression (SSA-RFR) models with excellent model applicability and accuracy. The temporal–spatial variations in the SOCS were modeled using 1080 soil samples and a set of nine environmental covariates. We analyzed the effect of land use changes on the SOCS in the WSP. The total SOCS increased by 18.03 Tg C from 1990 to 2020. The results of this study confirmed a significant increase in the SOCS in the study area since 2010. There was a 27.88 Tg C increase in the SOCS in 2020 compared to the total SOCS in 2010. We found that the spatial distribution of the SOCS increased from southeast to northwest, and the vertical distribution of the SOCS in the study area decreased with increasing soil depth. Forests and grasslands are the main sources of SOCS the total SOCS in the forest and grassland accounted for 37.53 and 59.39% of the total soil organic carbon (SOC) pool in 2020, respectively. The expansion of the wetlands, forest, and grassland areas could increase the SOCS in the study area. A timely and accurate understanding of the dynamics of SOC is crucial for developing effective land management strategies to enhance carbon sequestration and mitigate land degradation. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The location of the study area. (<b>b</b>) The monthly mean precipitation and temperature data for the study area in 2020. (<b>c</b>) The land use data in 2020 and the spatial distribution of the SOC sampling sites. (<b>d</b>) The map of the soil types in the study area.</p>
Full article ">Figure 2
<p>Schematic diagram of sampling area. (<b>a</b>) Distribution of samples in each sampling area. (<b>b</b>) Schematic diagram of soil profile.</p>
Full article ">Figure 3
<p>Workflow chart of this study.</p>
Full article ">Figure 4
<p>SOC content in different soil depths.</p>
Full article ">Figure 5
<p>The linear relationships between the SBD and the SOC (<b>a</b>), pH, and SOC (<b>b</b>).</p>
Full article ">Figure 6
<p>The SOCD of different land use types in the 0–30 cm interval. The white dots indicate the median and the upper and lower lines denote the maximum and minimum values of the SOCD.</p>
Full article ">Figure 7
<p>The importance of indicators in the SSA-RFR models for estimating the (<b>a</b>) SBD and (<b>b</b>) SOC in the different soil layers.</p>
Full article ">Figure 8
<p>Validation of simulated data using measured data. (<b>a</b>) Validation of SBD. (<b>b</b>) Validation of SOC.</p>
Full article ">Figure 9
<p>The spatial distribution of the SOCD at different soil depths in the WSP obtained using the SSA-RFR models.</p>
Full article ">Figure 10
<p>Spatial distribution of SOCD from 1990 to 2020.</p>
Full article ">Figure 11
<p>LUC in the study area from 1990 to 2020.</p>
Full article ">
12 pages, 5145 KiB  
Technical Note
Instrumentation for Sub-Ampere Lightning Current Measurement on a Tall Meteorological Tower in Complex Electromagnetic Environment
by Shaoyang Wang, Yan Gao, Mingli Chen, Zongxu Qiu, Hongbo Zhuang and Runquan Huang
Remote Sens. 2024, 16(7), 1307; https://doi.org/10.3390/rs16071307 - 8 Apr 2024
Viewed by 872
Abstract
Measurement of lightning current plays a critical role in the field of atmospheric electricity. Traditionally, Rogowski coils or low-resistance shunts were employed for measuring lightning currents in the range from several amperes up to several hundreds of kilo-amperes, and high-value resistors were utilized [...] Read more.
Measurement of lightning current plays a critical role in the field of atmospheric electricity. Traditionally, Rogowski coils or low-resistance shunts were employed for measuring lightning currents in the range from several amperes up to several hundreds of kilo-amperes, and high-value resistors were utilized for measuring corona discharge currents at sub-ampere levels. However, these approaches were not suitable for continuously recording the vast range of lightning currents. For this sake, we have developed a lightning current measurement system equipped with a shock-tolerant low-noise amplifier module. With the system installed on a tall tower, sub-ampere level currents just before the lightning initiation were observed for the first time. To confirm the authenticity of the recorded currents, the background noise of the measurement system and surrounding environment were identified, and a digital multi-frequency notch filter was proposed for de-noising. Results show that the system can achieve a current identification level of 50 mA even in complex electromagnetic environments, while having a measurement capability of 220 kA. Full article
Show Figures

Figure 1

Figure 1
<p>A block diagram for the implementation of the lightning current acquisition system and the isolated power supply (IPS) installed on and around the SZMGT, where the amplifier module indicated by a red box is recently implemented and specially designed for measuring the lightning corona discharge current pulses.</p>
Full article ">Figure 2
<p>The internal configuration of the lightning current measurement enclosure installed on the upper platform of the SZMGT.</p>
Full article ">Figure 3
<p>The frequency responses of the high-gain amplifier tested on 25 July 2018 (just after the assembly) and 10 August 2020 (after the shock testing and two lightning events).</p>
Full article ">Figure 4
<p>The noise performance of the high-gain amplifier tested on 25 July 2018 (just after the assembly) and 10 August 2020 (after the shock testing and two lightning events).</p>
Full article ">Figure 5
<p>Smoothed RTI noise spectral densities of the bypass channel and the AFE high-gain output channel under different E/O settings. It should be noted that since the bandwidth setting of E/O has little effect on the noise, only the noise performance with internal 6th order 10 MHz Bessel low-pass filter is drawn for simplicity.</p>
Full article ">Figure 6
<p>RTI integral noises of the bypass channel and the AFE high-gain output channel under different E/O settings.</p>
Full article ">Figure 7
<p>Comparisons of the noise spectrum measured on a sunny day (<b>a</b>) with the signal spectrum of the observed lightning current waveform (<b>b</b>), and the identification thresholds for the frequencies of the inserted notch filter (red line).</p>
Full article ">Figure 8
<p>Observed lightning current waveforms on 16 September 2019, where panel (<b>a</b>) shows the whole event and panels (<b>b</b>–<b>d</b>) provide zoomed-in views of the lavender box in the panel above it. Ch1 is for the bypass channel and Ch2 is for the high-gain channel.</p>
Full article ">Figure 9
<p>Observed lightning current waveforms on 30 May 2020, where the panel (<b>a</b>) shows the whole event and panels (<b>b</b>–<b>d</b>) provide zoomed-in views of the lavender box in the panel above it.</p>
Full article ">
23 pages, 33239 KiB  
Article
Lunar Surface Resource Exploration: Tracing Lithium, 7 Li and Black Ice Using Spectral Libraries and Apollo Mission Samples
by Susana del Carmen Fernández, Fernando Alberquilla, Julia María Fernández, Enrique Díez, Javier Rodríguez, Rubén Muñiz, Javier F. Calleja, Francisco Javier de Cos and Jesús Martínez-Frías
Remote Sens. 2024, 16(7), 1306; https://doi.org/10.3390/rs16071306 - 8 Apr 2024
Viewed by 1144
Abstract
This is an exercise to explore the concentration of lithium, lithium-7 isotope and the possible presence of black dirty ice on the lunar surface using spectral data obtained from the Clementine mission. The main interest in tracing the lithium and presence of dark [...] Read more.
This is an exercise to explore the concentration of lithium, lithium-7 isotope and the possible presence of black dirty ice on the lunar surface using spectral data obtained from the Clementine mission. The main interest in tracing the lithium and presence of dark ice on the lunar surface is closely related to future human settlement missions on the moon. We investigate the distribution of lithium and 7 Li isotope on the lunar surface by employing spectral data from the Clementine images. We utilized visible (VIS–NIR) imagery at wavelengths of 450, 750, 900, 950 and 1000 nm, along with near-infrared (NIR–SWIR) at 1100, 1250, 1500, 2000, 2600 and 2780 nm, encompassing 11 bands in total. This dataset offers a comprehensive coverage of about 80% of the lunar surface, with resolutions ranging from 100 to 500 m, spanning latitudes from 80°S to 80°N. In order to extract quantitative abundance of lithium, ground-truth sites were used to calibrate the Clementine images. Samples (specifically, 12045, 15058, 15475, 15555, 62255, 70035, 74220 and 75075) returned from Apollo missions 12, 15, 16 and 17 have been correlated to the Clementine VIS–NIR bands and five spectral ratios. The five spectral ratios calculated synthesize the main spectral features of sample spectra that were grouped by their lithium and 7 Li content using Principal Component Analysis. The ratios spectrally characterize substrates of anorthosite, silica-rich basalts, olivine-rich basalts, high-Ti mare basalts and Orange and Glasses soils. Our findings reveal a strong linear correlation between the spectral parameters and the lithium content in the eight Apollo samples. With the values of the 11 Clementine bands and the 5 spectral ratios, we performed linear regression models to estimate the concentration of lithium and 7 Li. Also, we calculated Digital Terrain Models (Altitude, Slope, Aspect, DirectInsolation and WindExposition) from LOLA-DTM to discover relations between relief and spatial distribution of the extended models of lithium and 7 Li. The analysis was conducted in a mask polygon around the Apollo 15 landing site. This analysis seeks to uncover potential 7 Li enrichment through spallation processes, influenced by varying exposure to solar wind. To explore the possibility of finding ice mixed with regolith (often referred to as `black ice’), we extended results to the entire Clementine coverage spectral indices, calculated with a library (350–2500 nm) of ice samples contaminated with various concentrations of volcanic particles. Full article
(This article belongs to the Special Issue Future of Lunar Exploration)
Show Figures

Figure 1

Figure 1
<p>Cartographic expression of the Wind Exposition Index. The areas in dark grey represent those less exposed to solar wind, while the areas in light grey represent those experiencing a greater impact from solar wind.</p>
Full article ">Figure 2
<p>Training area (around the Apollo 15 landing site) and point population used to test the influence of relief variables in lithium distribution.</p>
Full article ">Figure 3
<p>On the left, taking the spectral library (VIS–NIR) over Johnson Glacier; in the middle, the Black Glacier of Deception Inland; on the right, front of Johnson Glacier, covered with lapilli.</p>
Full article ">Figure 4
<p>On the left, extraction and characterization in the Juan Carlos I Spanish Scientific Base in Livingstone Island of impurities from a snow sample of the Johnsons Glacier. On the right, we can see the ice rendered dirty by us with lapilli. In the middle, the ADS spectroradiometer taking spectra of dirty ice at temperatures below 0 °C.</p>
Full article ">Figure 5
<p>Types of samples in the spectral library of ices (0, 2, 4 and 6 ppm of very fine, &lt;2 mm, lapilli with andesitic composition). In the case of dirty ices, we do not have “ground truth” to compare or validate the results. The four ice ratios were regionalized to the entire Clementine image and we then visualized the results.</p>
Full article ">Figure 6
<p>Correlation between exposure age (38Ar) and 7 Li concentration in lunar samples without 62255 and 74220.</p>
Full article ">Figure 7
<p>Relations between altitude (m) of samples and concentration of 7 Li.</p>
Full article ">Figure 8
<p>Five spectra types of lunar samples obtained in PCA analysis.</p>
Full article ">Figure 9
<p>Detailed cartographic expression of C5 (Orange and Glasses soils) spectral index crossing with the Apollo sample 74220.</p>
Full article ">Figure 10
<p>Extended lithium (ppm) and 7 Li‰ estimation models. Detailed mapping in the area of the Apollo 15 landing site. The base maps are Clementine image (500 m/pixel) and LOLA (50 m/pixel). The projection system is GCS_moon_2000.</p>
Full article ">Figure 11
<p>Attempt to validate the extended models with the eight lunar samples.</p>
Full article ">Figure 12
<p>Correlations between values of relief variables derived from DEM_LOLA and lithium and 7 Li concentrations. The analysis were performed with a simulated population of 296 samples concentrated in the area of Apollo 15 landing site.</p>
Full article ">Figure 13
<p>ADS Vis_Nir spectral library for ices.</p>
Full article ">Figure 14
<p>Cartographic expression of the spectral ice indices (ICELx1, ICELx2 and ICELx3) across the entire Clementine image. The frequencies of the values obtained by each of the indices in the Clementine image can be consulted in the corresponding histograms.</p>
Full article ">Figure 15
<p>Example of regolith with ice probability distribution.</p>
Full article ">
26 pages, 49819 KiB  
Article
Comparison of Cloud-Mask Algorithms and Machine-Learning Methods Using Sentinel-2 Imagery for Mapping Paddy Rice in Jianghan Plain
by Xinyi Gao, Hong Chi, Jinliang Huang, Yifei Han, Yifan Li and Feng Ling
Remote Sens. 2024, 16(7), 1305; https://doi.org/10.3390/rs16071305 - 8 Apr 2024
Viewed by 1269
Abstract
Southern China, one of the traditional rice production bases, has experienced significant declines in the area of rice paddy since the beginning of this century. Monitoring the rice cropping area is becoming an urgent need for food security policy decisions. One of the [...] Read more.
Southern China, one of the traditional rice production bases, has experienced significant declines in the area of rice paddy since the beginning of this century. Monitoring the rice cropping area is becoming an urgent need for food security policy decisions. One of the main challenges for mapping rice in this area is the quantity of cloud-free observations that are vulnerable to frequent cloud cover. Another relevant issue that needs to be addressed is determining how to select the appropriate classifier for mapping paddy rice based on the cloud-masked observations. Therefore, this study was organized to quickly find a strategy for rice mapping by evaluating cloud-mask algorithms and machine-learning methods for Sentinel-2 imagery. Specifically, we compared four GEE-embedded cloud-mask algorithms (QA60, S2cloudless, CloudScore, and CDI (Cloud Displacement Index)) and analyzed the appropriateness of widely accepted machine-learning classifiers (random forest, support vector machine, classification and regression tree, gradient tree boost) for cloud-masked imagery. The S2cloudless algorithm had a clear edge over the other three algorithms based on its overall accuracy in evaluation and visual inspection. The findings showed that the algorithm with a combination of S2cloudless and random forest showed the best performance when comparing mapping results with field survey data, referenced rice maps, and statistical yearbooks. In general, the research highlighted the potential of using Sentinel-2 imagery to map paddy rice with multiple combinations of cloud-mask algorithms and machine-learning methods in a cloud-prone area, which has the potential to broaden our rice mapping strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Overview map of the study area. The grids are Sentinel-2 footprints in MGRS (Military Grid Reference System) with an area of 100 km × 100 km square. Land cover data was from the optimal mapping results of this study.</p>
Full article ">Figure 2
<p>The workflow of the study included data preprocessing and cloud mask algorithm evaluation, sample selection and feature selection, extraction of rice phenology and image compositing, comparisons of machine-learning algorithms in rice mapping and validation of rice maps (validation of field data samples, comparison of 10 m rice maps, and comparison of statistical data).</p>
Full article ">Figure 3
<p>Paddy rice phenological stages derived from the fitted curves of five spectral indices based on the cloud-masked TOA dataset in 2021 (field photographs were taken at <math display="inline"><semantics> <mrow> <mn>112</mn> <mo>.</mo> <msup> <mn>8804</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>E, <math display="inline"><semantics> <mrow> <mn>30</mn> <mo>.</mo> <msup> <mn>0690</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>N).</p>
Full article ">Figure 4
<p>Results of four cloud-mask algorithms in the tile of 49REP, 49RFP, and 49RGP (the specific locations of these footprints in the study area are shown in <a href="#remotesensing-16-01305-f001" class="html-fig">Figure 1</a>). Each row shows the cloud identification results based on 49REP’s May 2018 TOA data, 49RFP’s July 2021 TOA data, and 49RGP’s September 2021 SR data using QA60, S2cloudless, CloudScore, and CDI cloud mask algorithms, respectively.</p>
Full article ">Figure 5
<p>Results of four cloud-mask algorithms in three subregions across the tiles of 49REP, 49RFP, and 49RGP with different land cover characteristics. Panels a, b, and c show the results in a region mixed with built-up area and paddy rice, a region mixed with dryland and paddy rice, and a region mixed with paddy rice and aquaculture area, respectively.</p>
Full article ">Figure 6
<p>J-M values of paddy rice and other land cover types to cloud-free datasets for different cloud-mask algorithms.</p>
Full article ">Figure 7
<p>Percentages of the specified land cover area to the total area in different land cover types were estimated from the different combinations of cloud-mask algorithms and machine-learning algorithms. The bars show the percentage in area of water body, built-up area, forest land, dryland, and paddy rice from left to right, respectively.</p>
Full article ">Figure 8
<p>J-sim values between the reference rice maps and the maps generated from different algorithm combinations.</p>
Full article ">Figure 9
<p>Comparisons of rice area between statistical data and mapping results.</p>
Full article ">Figure A1
<p>Rice distribution results based on 2018 TOA data.</p>
Full article ">Figure A2
<p>Rice distribution results based on 2021 TOA data.</p>
Full article ">Figure A3
<p>Rice distribution results based on 2021 SR data.</p>
Full article ">
20 pages, 52640 KiB  
Article
Weighted Differential Gradient Method for Filling Pits in Light Detection and Ranging (LiDAR) Canopy Height Model
by Guoqing Zhou, Haowen Li, Jing Huang, Ertao Gao, Tianyi Song, Xiaoting Han, Shuaiguang Zhu and Jun Liu
Remote Sens. 2024, 16(7), 1304; https://doi.org/10.3390/rs16071304 - 8 Apr 2024
Cited by 1 | Viewed by 956
Abstract
The canopy height model (CHM) derived from LiDAR point cloud data is usually used to accurately identify the position and the canopy dimension of single tree. However, local invalid values (also called data pits) are often encountered during the generation of CHM, which [...] Read more.
The canopy height model (CHM) derived from LiDAR point cloud data is usually used to accurately identify the position and the canopy dimension of single tree. However, local invalid values (also called data pits) are often encountered during the generation of CHM, which results in low-quality CHM and failure in the detection of treetops. For this reason, this paper proposes an innovative method, called “pixels weighted differential gradient”, to filter these data pits accurately and improve the quality of CHM. First, two characteristic parameters, gradient index (GI) and Z-score value (ZV) are extracted from the weighted differential gradient between the pit pixels and their eight neighbors, and then GIs and ZVs are commonly used as criterion for initial identification of data pits. Secondly, CHMs of different resolutions are merged, using the image processing algorithm developed in this paper to distinguish either canopy gaps or data pits. Finally, potential pits were filtered and filled with a reasonable value. The experimental validation and comparative analysis were carried out in a coniferous forest located in Triangle Lake, United States. The experimental results showed that our method could accurately identify potential data pits and retain the canopy structure information in CHM. The root-mean-squared error (RMSE) and mean bias error (MBE) from our method are reduced by between 73% and 26% and 76% and 28%, respectively, when compared with six other methods, including the mean filter, Gaussian filter, median filter, pit-free, spike-free and graph-based progressive morphological filtering (GPMF). The average F1 score from our method could be improved by approximately 4% to 25% when applied in single-tree extraction. Full article
Show Figures

Figure 1

Figure 1
<p>The flowchart of the proposed method.</p>
Full article ">Figure 2
<p>A pit pixel with its eight neighborhoods: (<b>a</b>) canopy of a tree and (<b>b</b>) a pit pixel.</p>
Full article ">Figure 3
<p>Canopy height model with a resolution of 0.5 m: (<b>a</b>) pits and (<b>b</b>) canopy gaps.</p>
Full article ">Figure 4
<p>Morphologic closure operation for raw CHM image: (<b>a</b>) raw CHM; (<b>b</b>) processed CHM.</p>
Full article ">Figure 5
<p>Study area located in Triangle Lake, Oregon, USA: (<b>a</b>) map of study area; (<b>b</b>) study area coverage; and (<b>c</b>) the three selected test areas.</p>
Full article ">Figure 6
<p>Three test areas with trees of different heights and stem densities: (<b>a</b>) is the test area 1, (<b>b</b>) represents the test area 2, and (<b>c</b>) represents test area 3.</p>
Full article ">Figure 7
<p>The LiDAR point cloud data preprocessing: (<b>a</b>) raw LiDAR point cloud data, (<b>b</b>) classified point cloud data, and (<b>c</b>) normalized point cloud data.</p>
Full article ">Figure 8
<p>CHM at different spatial resolutions: (<b>a</b>) is with 2.0 m spatial resolution; (<b>b</b>) is 1.0 m spatial resolution; (<b>c</b>) is with 0.5 m spatial resolution; and (<b>d</b>) is with 0.25 m spatial resolution.</p>
Full article ">Figure 9
<p>Extraction of characteristic parameters: (<b>a</b>) CHM, (<b>b</b>) GI, and (<b>c</b>) ZV; (<b>a1</b>–<b>c1</b>) zoom display; (<b>a2</b>–<b>c2</b>) pixel values in the window.</p>
Full article ">Figure 10
<p>(<b>a</b>) The merging CHMs with different spatial resolutions, where (<b>a</b>) represents a merged images at 2.0 m (<b>a1</b>), 1.0 m (<b>a2</b>), 0.5 m (<b>a3</b>), 0.25 m (<b>a4</b>); (<b>b</b>) represents the distinguished canopy pit pixels.</p>
Full article ">Figure 11
<p>The CHMs of test area 1: (<b>a</b>) before pit filling and (<b>b</b>) after pit filling.</p>
Full article ">Figure 12
<p>Extraction of characteristic parameters for test area 2. (<b>a</b>) CHM, (<b>b</b>) GI, and (<b>c</b>) ZV.</p>
Full article ">Figure 13
<p>The extracted characteristic parameters for test area 3. (<b>a</b>) CHM, (<b>b</b>) GI, and (<b>c</b>) ZV.</p>
Full article ">Figure 14
<p>Merging CHMs with different spatial resolutions: (<b>a</b>) represents a merged image at 2.0 m (<b>a1</b>), 1.0 m (<b>a2</b>), 0.5 m (<b>a3</b>), 0.25 m (<b>a4</b>); (<b>b</b>) represents the distinguished canopy pit pixels.</p>
Full article ">Figure 15
<p>Merging CHMs with different spatial resolutions: (<b>a</b>) represents a merged images at 2.0 m (<b>a1</b>), 1.0 m (<b>a2</b>), 0.5 m (<b>a3</b>), 0.25 m (<b>a4</b>); (<b>b</b>) represents the distinguished canopy pit pixels.</p>
Full article ">Figure 16
<p>The CHMs of test area 2 with (<b>a</b>) before pit filling and (<b>b</b>) after pit filling.</p>
Full article ">Figure 17
<p>The CHMs of test area 3 with (<b>a</b>) before pit filling and (<b>b</b>) after pit filling.</p>
Full article ">Figure 18
<p>Comparison analysis between the CHMs with 0.5 m resolution created by six pit-filling algorithms in test area 1: (<b>a</b>) is the original CHM, (<b>b</b>) is from the mean filter method, (<b>c</b>) is from the Gaussian filter, (<b>d</b>) is from the median filter, (<b>e</b>) is from the pit-free algorithm, (<b>f</b>) is from the spike-free algorithm, (<b>g</b>) is from the GPMF algorithm, and (<b>h</b>) is from the method proposed in this paper. (<b>a1</b>–<b>h1</b>) zoom display.</p>
Full article ">Figure 19
<p>Comparison analysis between the CHMs with 0.5 m resolution created by six pit-filling algorithms in test area 2: (<b>a</b>) is the original CHM, (<b>b</b>) is from the mean filter method, (<b>c</b>) is from the Gaussian filter, (<b>d</b>) is from the median filter, (<b>e</b>) is from the pit-free algorithm, (<b>f</b>) is from the spike-free algorithm, (<b>g</b>) is from the GPMF algorithm, and (<b>h</b>) is from the method proposed in this paper. (<b>a1</b>–<b>h1</b>) zoom display.</p>
Full article ">Figure 20
<p>Comparison analysis between the CHMs with 0.5 m resolution created by six pit-filling algorithms in test area 3: (<b>a</b>) is the original CHM, (<b>b</b>) is from the mean filter method, (<b>c</b>) is from the Gaussian filter, (<b>d</b>) is from the median filter, (<b>e</b>) is from the pit-free algorithm, (<b>f</b>) is from the spike-free algorithm, (<b>g</b>) is from the GPMF algorithm, and (<b>h</b>) is from the method proposed in this paper. (<b>a1</b>–<b>h1</b>) zoom display.</p>
Full article ">Figure 21
<p>The profile for comparison analysis of six methods with a localized single tree at the three test areas: (<b>a</b>) test area 1, (<b>b</b>) test area 2, (<b>c</b>) test area 3.</p>
Full article ">Figure 22
<p>Comparison between the CHM in mixed forest: (<b>a</b>) the raw CHM and (<b>b</b>) the filled CHM using our method: (<b>a1</b>,<b>a2</b>) zoom display for raw CHM, (<b>b1</b>,<b>b2</b>) zoom display for filled CHM.</p>
Full article ">
22 pages, 5599 KiB  
Article
Landslide-Hazard-Avoiding Highway Alignment Selection in Mountainous Regions Based on SAR Images and High-Spatial-Resolution Precipitation Datasets: A Case Study in Southwestern China
by Zhiheng Wang, Yang Jia, Shengfu Li, Rui Zhang, Binzhi Xu and Xiaopeng Sun
Remote Sens. 2024, 16(7), 1303; https://doi.org/10.3390/rs16071303 - 8 Apr 2024
Cited by 1 | Viewed by 980
Abstract
Landslides recurrently cause severe damage and, in some cases, the full disruption of many highways in mountainous areas, which can last from a few days to even months. Thus, there is a high demand for monitoring tools and precipitation data to support highway [...] Read more.
Landslides recurrently cause severe damage and, in some cases, the full disruption of many highways in mountainous areas, which can last from a few days to even months. Thus, there is a high demand for monitoring tools and precipitation data to support highway alignment selections before construction. In this study, we proposed a new system highway alignment selection method based on coherent scatter InSAR (CSI) and ~1 km high-spatial-resolution precipitation (HSRP) analysis. Prior to the CSI, we calculated and analyzed the feasibility of Sentinel-1A ascending and descending data. To illustrate the performance of the CSI, CSI and SBAS–InSAR were both utilized to monitor 80 slow-moving landslides, which were identified by optical remote-sensing interpretation and field investigation, along the Barkam–Kangting Highway Corridor (BKHC) in southwestern China, relying on 56 Sentinel-1A descending images from September 2019 to September 2021. The results reveal that CSI has clearer deformation signals and more measurement points (MPs) than SBAS-InSAR. And the maximum cumulative displacements and rates of the landslides reach −75 mm and −64 mm/year within the monitoring period (CSI results), respectively. Furthermore, the rates of the landslides near the Jinchuan River are higher than those of the landslides far from the river. Subsequently, to optimize the highway alignment selection, we analyzed the spatiotemporal evolution characteristics of feature points on a typical landslide by combining the −1 km HSRP, which was calculated from the 30′ Climatic Research Unit (CRU) time-series datasets, with the climatology datasets of WorldClim using delta spatial downscaling. The analysis shows that the sliding rates of landslides augment from the back edge to the tongue because of fluvial erosion and that accelerated sliding is highly related to the intense precipitation between April and September each year (ASP). Consequently, three solution types were established in our method by setting thresholds for the deformation rates and ASPs of every landslide. Afterward, the risk-optimal alignment selection of the BKHC was finalized according to the solution types and consideration of the construction’s possible impacts. Ultimately, the major problems and challenges for our method were discussed, and conclusions were given. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Figure 1

Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>Faults in the study area.</p>
Full article ">Figure 3
<p>Distribution of national weather stations across China (modified from Figure 1 from in [<a href="#B62-remotesensing-16-01303" class="html-bibr">62</a>]).</p>
Full article ">Figure 4
<p>Method flow.</p>
Full article ">Figure 5
<p>Relationship between the LOS and the downslope displacements for different slope orientations (scenarios 1, 2, and 3 are facing the sensor and scenarios 4, 5, and 6 are facing away from the sensor) and slopes (α and α’) (adapted with permission from Ref. [<a href="#B67-remotesensing-16-01303" class="html-bibr">67</a>]. 2016, Keren Dai).</p>
Full article ">Figure 6
<p>The FSs of the Sentinel-1A data: (<b>a</b>) ascending and (<b>b</b>) descending; (<b>c</b>) the terrain of the study area.</p>
Full article ">Figure 7
<p>The LOS deformation rates derived using SBAS-InSAR and CSI: (<b>a</b>) SBAS-InSAR and (<b>b</b>) CSI; (<b>c</b>) ~1 km HSRP in June 2020.</p>
Full article ">Figure 8
<p>The annual deformation rates derived from descending Sentinel-1a data using two TS–InSAR methods: (<b>a</b>) SBAS-InSAR and (<b>b</b>) CSI; (<b>c</b>) is the optical remote-sensing interpretation; (<b>d</b>) is the field investigation.</p>
Full article ">Figure 9
<p>The cumulative displacement curves of typical points and monthly HSRPs of the landslide.</p>
Full article ">Figure 10
<p>Profile map of the landslide and the profile line along A1–A2.</p>
Full article ">
13 pages, 8931 KiB  
Article
Early Identification of River Blockage Disasters Caused by Debris Flows in the Bailong River Basin, China
by Jianjun Zeng, Yan Zhao, Jiaoyu Zheng, Yongjun Zhang, Pengqing Shi, Yajun Li, Guan Chen, Xingmin Meng and Dongxia Yue
Remote Sens. 2024, 16(7), 1302; https://doi.org/10.3390/rs16071302 - 7 Apr 2024
Cited by 1 | Viewed by 1218
Abstract
The Bailong River Basin is one of the most developed regions for debris flow disasters worldwide, often causing severe secondary disasters by blocking rivers. Therefore, the early identification of potential debris flow disasters that may block the river in this region is of [...] Read more.
The Bailong River Basin is one of the most developed regions for debris flow disasters worldwide, often causing severe secondary disasters by blocking rivers. Therefore, the early identification of potential debris flow disasters that may block the river in this region is of great significance for disaster risk prevention and reduction. However, it is quite challenging to identify potential debris flow disasters that may block rivers at a regional scale, as conducting numerical simulations for each debris flow catchment would require significant time and financial resources. The purpose of this article is to use public resource data and machine learning methods to establish a relationship model between debris flow-induced river blockage and key influencing factors, thereby economically predicting potential areas at risk for debris flow-induced river blockage disasters. Based on the field investigation, data collection, and remote sensing interpretation, this study selected 12 parameters, including the basin area, basin height difference, relief ratio, circularity ratio, landslide density, fault density, lithology index, annual average frequency of daily rainfall exceeding 40 mm, river width, river discharge, river gradient, and confluence angle, as critical factors to determine whether debris flows will cause river blockages. A relationship model between debris flow-induced river blockage and influencing factors was constructed based on machine learning algorithms. Several machine learning algorithms were compared, and the XGB model performed the best, with a prediction accuracy of 0.881 and an area under the ROC curve of 0.926. This study found that the river width is the determining factor for debris flow blocking rivers, followed by the annual average frequency of daily rainfall exceeding 40 mm, basin height difference, circularity ratio, basin area, and river discharge. The early identification method proposed in this study for river blockage disasters caused by debris flows can provide a reference for the quantitative assessment and pre-disaster prevention of debris flow-induced river blockage chain risks in similar high-mountain gorge areas. Full article
Show Figures

Figure 1

Figure 1
<p>The distribution of debris flow catchments once blocked the river and catchments without river blocking records in the Bailong River Basin (within Gansu Province, China).</p>
Full article ">Figure 2
<p>The distribution maps of the landslide density (<b>a</b>), fault density (<b>b</b>), lithologic index (<b>c</b>), and annual average frequency of daily rainfall exceeding 40 mm (<b>d</b>) in the Bailong River Basin.</p>
Full article ">Figure 3
<p>Correlation heatmap of influencing factors.</p>
Full article ">Figure 4
<p>The prediction accuracy of the model in the validation set samples.</p>
Full article ">Figure 5
<p>ROC curve of XGB model.</p>
Full article ">Figure 6
<p>Identification results of debris flow river-blocking disasters in Bailong River Basin.</p>
Full article ">Figure 7
<p>Feature importance of different factors influencing river blockage.</p>
Full article ">Figure 8
<p>Box diagrams of important parameters.</p>
Full article ">
15 pages, 3942 KiB  
Technical Note
GPU Acceleration for SAR Satellite Image Ortho-Rectification
by Lei Dong, Tingtao Zhang, Fangjian Liu, Rui Liu and Hongjian You
Remote Sens. 2024, 16(7), 1301; https://doi.org/10.3390/rs16071301 - 7 Apr 2024
Cited by 1 | Viewed by 1089
Abstract
Synthetic Aperture Radar (SAR) satellite image ortho-rectification requires pixel-level calculations, which are time-consuming. Moreover, for SAR images with large overlapping areas, the processing time for ortho-rectification increases linearly, significantly reducing the efficiency of SAR satellite image mosaic. This paper thoroughly analyzes two geometric [...] Read more.
Synthetic Aperture Radar (SAR) satellite image ortho-rectification requires pixel-level calculations, which are time-consuming. Moreover, for SAR images with large overlapping areas, the processing time for ortho-rectification increases linearly, significantly reducing the efficiency of SAR satellite image mosaic. This paper thoroughly analyzes two geometric positioning models for SAR images. In order to address the high computation time of pixel-by-pixel ortho-rectification in SAR satellite images, a GPU-accelerated pixel-by-pixel correction method based on a rational polynomial coefficients (RPCs) model is proposed, which improves the efficiency of SAR satellite image ortho-rectification. Furthermore, in order to solve the problem of linearly increasing processing time for the ortho-rectification of multiple SAR images in large overlapping areas, a multi-GPU collaborative acceleration strategy for the ortho-rectification of multiple SAR images in large overlapping areas is proposed, achieving efficient ortho-rectification processing of multiple SAR image data in large overlapping areas. By conducting ortho-rectification experiments on 20 high-resolution SAR images from the Gaofen-3 satellite, the feasibility and efficiency of the multi-GPU collaborative acceleration processing algorithm are verified. Full article
(This article belongs to the Special Issue Calibration and Validation of SAR Data and Derived Products)
Show Figures

Figure 1

Figure 1
<p>GPU-accelerated pixel-by-pixel ortho-rectification of SAR images based on the RPC rational function model.</p>
Full article ">Figure 2
<p>Distribution of original experimental data.</p>
Full article ">Figure 3
<p>Orthorectified image results.</p>
Full article ">Figure 4
<p>Parallel speed-up ratio of SAR imagery ortho-rectification.</p>
Full article ">Figure 5
<p>Parallel performance improvement ratio of SAR imagery ortho-rectification.</p>
Full article ">Figure 6
<p>Comparison of ortho-rectification time for SAR images.</p>
Full article ">
11 pages, 3477 KiB  
Communication
Separation of Rapidly-Varying and Slowly-Varying Processes and Development of Diffraction Decomposition Order Method in Radiative Transfer
by Meng Zhang, Chenxu Gao, Bingqiang Sun and Yijun Zhang
Remote Sens. 2024, 16(7), 1300; https://doi.org/10.3390/rs16071300 - 7 Apr 2024
Viewed by 887
Abstract
Single scattering in radiative transfer is separated into rapidly-varying and slowly-varying processes, where the rapidly-varying process (RVP) is mainly contributed by the diffraction effect. Accordingly, the diffraction decomposition order (DDO) method is developed to solve the vector radiative transfer equation (VRTE). Instead of [...] Read more.
Single scattering in radiative transfer is separated into rapidly-varying and slowly-varying processes, where the rapidly-varying process (RVP) is mainly contributed by the diffraction effect. Accordingly, the diffraction decomposition order (DDO) method is developed to solve the vector radiative transfer equation (VRTE). Instead of directly solving the original VRTE, we decompose it into a series of order equations, where the zeroth-order equation replaces the RVP with a δ-function while the high-order equations are the same as the zeroth-order one, except that the high decomposition orders of the RVP are used as driven sources. In this study, the DDO method is numerically realized using the successive order of the scattering method. The DDO is computationally efficient and accurate. More importantly, all physical processes in the VRTE are fully decomposed due to the order decomposition of the RVP and can be straightforwardly discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Exemplary scattering phase function as shown in the solid line and the corresponding accumulated scattering phase function as shown in the dashed line.</p>
Full article ">Figure 2
<p>The flow chart of the diffraction decomposition order (DDO) algorithm realized by the successive order of scattering (SOS) method. The signals S, M, and I denote single scattering, multiple scattering, and Stokes vector, respectively.</p>
Full article ">Figure 3
<p>Comparison of the reflected radiance I and relative differences in the case of a predefined aerosol layer. The solar zenith angle is 120° and the relative azimuthal angle is taken as 90°.</p>
Full article ">Figure 4
<p>Illustration of solar, upward and downward directions, and zenith and relative azimuthal angles in the numerical simulations.</p>
Full article ">Figure 5
<p>Original, rapidly-varying, and slowly-varying scattering phase matrix <math display="inline"><semantics> <mfenced separators="" open="(" close=")"> <mi mathvariant="bold">P</mi> <mo>,</mo> <msub> <mi mathvariant="bold">P</mi> <mi>r</mi> </msub> <mo>,</mo> <msub> <mi mathvariant="bold">P</mi> <mi>s</mi> </msub> </mfenced> </semantics></math> used in the following study. The predefined truncation angle for <math display="inline"><semantics> <msub> <mi mathvariant="bold">P</mi> <mi>r</mi> </msub> </semantics></math> is <math display="inline"><semantics> <mrow> <mn>4</mn> <mo>.</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>0.452</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Stokes parameters (I, Q) and their relative difference compared to the results calculated by the SOS method at layer top (<math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>, (<b>a</b>–<b>d</b>)), middle (<math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>2.5</mn> </mrow> </semantics></math>, (<b>e</b>–<b>h</b>)), and bottom (<math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mi>b</mi> </mrow> </semantics></math>, (<b>i</b>–<b>l</b>)) calculated by the straightforward SOS and the DDO with different orders.</p>
Full article ">Figure 7
<p>Stokes vector I (<b>a</b>) and the relative values of Q, U, V (<b>b</b>–<b>d</b>) and the relative differences (<b>e</b>,<b>f</b>) comparisons at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>2.5</mn> </mrow> </semantics></math> calculated by the SOS and the DDO methods with different orders when the VAA is <math display="inline"><semantics> <msup> <mn>60</mn> <mo>∘</mo> </msup> </semantics></math>.</p>
Full article ">Figure 8
<p>Absolute difference of Stokes parameters calculated by the straightforward SOS and the DDO with different orders at layer bottom.</p>
Full article ">
21 pages, 6228 KiB  
Article
An Improved SAR Ship Classification Method Using Text-to-Image Generation-Based Data Augmentation and Squeeze and Excitation
by Lu Wang, Yuhang Qi, P. Takis Mathiopoulos, Chunhui Zhao and Suleman Mazhar
Remote Sens. 2024, 16(7), 1299; https://doi.org/10.3390/rs16071299 - 7 Apr 2024
Cited by 3 | Viewed by 1288
Abstract
Synthetic aperture radar (SAR) plays a crucial role in maritime surveillance due to its capability for all-weather, all-day operation. However, SAR ship recognition faces challenges, primarily due to the imbalance and inadequacy of ship samples in publicly available datasets, along with the presence [...] Read more.
Synthetic aperture radar (SAR) plays a crucial role in maritime surveillance due to its capability for all-weather, all-day operation. However, SAR ship recognition faces challenges, primarily due to the imbalance and inadequacy of ship samples in publicly available datasets, along with the presence of numerous outliers. To address these issues, this paper proposes a SAR ship classification method based on text-generated images to tackle dataset imbalance. Firstly, an image generation module is introduced to augment SAR ship data. This method generates images from textual descriptions to overcome the problem of insufficient samples and the imbalance between ship categories. Secondly, given the limited information content in the black background of SAR ship images, the Tokens-to-Token Vision Transformer (T2T-ViT) is employed as the backbone network. This approach effectively combines local information on the basis of global modeling, facilitating the extraction of features from SAR images. Finally, a Squeeze-and-Excitation (SE) model is incorporated into the backbone network to enhance the network’s focus on essential features, thereby improving the model’s generalization ability. To assess the model’s effectiveness, extensive experiments were conducted on the OpenSARShip2.0 and FUSAR-Ship datasets. The performance evaluation results indicate that the proposed method achieves higher classification accuracy in the context of imbalanced datasets compared to eight existing methods. Full article
(This article belongs to the Special Issue Deep Learning Techniques Applied in Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>A structural comparison diagram between CNN and ViT.</p>
Full article ">Figure 2
<p>The pre-training process of CLIP.</p>
Full article ">Figure 3
<p>The overall framework of the proposed method.</p>
Full article ">Figure 4
<p>The operation of the diffusion model.</p>
Full article ">Figure 5
<p>The structure of the T2T module.</p>
Full article ">Figure 6
<p>The structure of the SE module.</p>
Full article ">Figure 7
<p>SAR ship samples from the OpenSARShip2.0 dataset representing (<b>a</b>) an optical image of Cargo; (<b>b</b>) a SAR image of Cargo; (<b>c</b>) an optical image of Fishing; (<b>d</b>) a SAR image of Fishing; (<b>e</b>) an optical image of Tug; and (<b>f</b>) a SAR image of Tug.</p>
Full article ">Figure 8
<p>SAR ship samples in FUSAR-Ship dataset. Among these, (<b>a</b>–<b>e</b>) represent optical images of Bulk Carrier, Cargo, Fishing, Tanker, and Other, and (<b>f</b>–<b>j</b>) represent SAR images of Bulk Carrier, Cargo, Fishing, Tanker, and Other.</p>
Full article ">Figure 9
<p>Image results generated by inputting text information: SAR Cargo, SAR Fishing, SAR Tug; (<b>a</b>–<b>c</b>) represent Cargo images, while (<b>d</b>–<b>f</b>) represent Fishing images, and (<b>g</b>–<b>i</b>) represent Tug images.</p>
Full article ">
0 pages, 18976 KiB  
Article
Spatiotemporal Patterns of Air Pollutants over the Epidemic Course: A National Study in China
by Kun Qin, Zhanpeng Wang, Shaoqing Dai, Yuchen Li, Manyao Li, Chen Li, Ge Qiu, Yuanyuan Shi, Chun Yin, Shujuan Yang and Peng Jia
Remote Sens. 2024, 16(7), 1298; https://doi.org/10.3390/rs16071298 - 7 Apr 2024
Cited by 1 | Viewed by 1457
Abstract
Air pollution has been standing as one of the most pressing global challenges. The changing patterns of air pollutants at different spatial and temporal scales have been substantially studied all over the world, which, however, were intricately disturbed by COVID-19 and subsequent containment [...] Read more.
Air pollution has been standing as one of the most pressing global challenges. The changing patterns of air pollutants at different spatial and temporal scales have been substantially studied all over the world, which, however, were intricately disturbed by COVID-19 and subsequent containment measures. Understanding fine-scale changing patterns of air pollutants at different stages over the epidemic’s course is necessary for better identifying region-specific drivers of air pollution and preparing for environmental decision making during future epidemics. Taking China as an example, this study developed a multi-output LightGBM approach to estimate monthly concentrations of the six major air pollutants (i.e., PM2.5, PM10, NO2, SO2, O3, and CO) in China and revealed distinct spatiotemporal patterns for each pollutant over the epidemic’s course. The 5-year period of 2019–2023 was selected to observe changes in the concentrations of air pollutants from the pre-COVID-19 era to the lifting of all containment measures. The performance of our model, assessed by cross-validation R2, demonstrated high accuracy with values of 0.92 for PM2.5, 0.95 for PM10, 0.95 for O3, 0.90 for NO2, 0.79 for SO2, and 0.82 for CO. Notably, there was an improvement in the concentrations of particulate matter, particularly for PM2.5, although PM10 exhibited a rebound in northern regions. The concentrations of SO2 and CO consistently declined across the country over the epidemic’s course (p < 0.001 and p < 0.05, respectively), while O3 concentrations in southern regions experienced a notable increase. Concentrations of air pollutants in the Beijing–Tianjin–Hebei region were effectively controlled and mitigated. The findings of this study provide critical insights into changing trends of air quality during public health emergencies, help guide the development of targeted interventions, and inform policy making aimed at reducing disease burdens associated with air pollution. Full article
(This article belongs to the Special Issue Geographic Data Analysis and Modeling in Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of national automatic air quality monitoring stations in China.</p>
Full article ">Figure 2
<p>Flowchart of the modeling and analysis process for this study. LightGBM, light gradient boosting machine; MODIS, moderate resolution imaging spectroradiometer; TROPOMI, tropospheric monitoring instrument.</p>
Full article ">Figure 3
<p>Spatial distribution of annual mean concentration of PM<sub>2.5</sub> (<b>a</b>), PM<sub>10</sub> (<b>b</b>), NO<sub>2</sub> (<b>c</b>), SO<sub>2</sub> (<b>d</b>), O<sub>3</sub> (<b>e</b>), and CO (<b>f</b>) in China from 2019 to 2023. The unit is mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 3 Cont.
<p>Spatial distribution of annual mean concentration of PM<sub>2.5</sub> (<b>a</b>), PM<sub>10</sub> (<b>b</b>), NO<sub>2</sub> (<b>c</b>), SO<sub>2</sub> (<b>d</b>), O<sub>3</sub> (<b>e</b>), and CO (<b>f</b>) in China from 2019 to 2023. The unit is mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 4
<p>Monthly mean concentrations of the six major air pollutants (<b>a</b>) and their interannual differences (<b>b</b>) in China from 2019 to 2023. The unit is mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 5
<p>Spatial distribution patterns (<b>left</b>) and the corresponding temporal trends (<b>right</b>) of air pollutants in China from 2019 to 2023, with <span class="html-italic">p</span>-values of the significant trends marked.</p>
Full article ">Figure 6
<p>Importance (%) of each feature during model construction.</p>
Full article ">Figure 7
<p>Density scatter plots of 10-fold cross-validation (CV) results of our multi-output LightGBM model. Solid lines denote the best-fit lines derived from linear regression, and dashed lines denote the 1:1 line. The provided information includes the sample size (N), coefficient of determination (R<sup>2</sup>), root-mean-square error (RMSE), and mean absolute error (MAE). The units of the RMSE and MAE are mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 7 Cont.
<p>Density scatter plots of 10-fold cross-validation (CV) results of our multi-output LightGBM model. Solid lines denote the best-fit lines derived from linear regression, and dashed lines denote the 1:1 line. The provided information includes the sample size (N), coefficient of determination (R<sup>2</sup>), root-mean-square error (RMSE), and mean absolute error (MAE). The units of the RMSE and MAE are mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 8
<p>Spatial distributions of the site-based cross-validation results. RMSE, root-mean-square error. The units of the RMSE are mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants.</p>
Full article ">Figure 9
<p>Density scatter plots of yearly sample-based cross-validation (CV) results across China from 2019 to 2023. Solid lines denote the best-fit lines derived from linear regression, and dashed lines denote the 1:1 line. The provided information includes the sample size (N), coefficient of determination (R<sup>2</sup>), root-mean-square error (RMSE), and mean absolute error (MAE). The units of the RMSE and MAE are mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants. The pollutants from left to right are PM<sub>2.5</sub> (<b>a</b>), PM<sub>10</sub> (<b>b</b>), NO<sub>2</sub> (<b>c</b>), SO<sub>2</sub> (<b>d</b>), O<sub>3</sub> (<b>e</b>), and CO (<b>f</b>). Shown from top to bottom are the years 2019–2023 in order.</p>
Full article ">Figure 9 Cont.
<p>Density scatter plots of yearly sample-based cross-validation (CV) results across China from 2019 to 2023. Solid lines denote the best-fit lines derived from linear regression, and dashed lines denote the 1:1 line. The provided information includes the sample size (N), coefficient of determination (R<sup>2</sup>), root-mean-square error (RMSE), and mean absolute error (MAE). The units of the RMSE and MAE are mg/m<sup>3</sup> for CO and µg/m<sup>3</sup> for other air pollutants. The pollutants from left to right are PM<sub>2.5</sub> (<b>a</b>), PM<sub>10</sub> (<b>b</b>), NO<sub>2</sub> (<b>c</b>), SO<sub>2</sub> (<b>d</b>), O<sub>3</sub> (<b>e</b>), and CO (<b>f</b>). Shown from top to bottom are the years 2019–2023 in order.</p>
Full article ">
18 pages, 4762 KiB  
Technical Note
SSEBop Evapotranspiration Estimates Using Synthetically Derived Landsat Data from the Continuous Change Detection and Classification Algorithm
by Mikael P. Hiestand, Heather J. Tollerud, Chris Funk, Gabriel B. Senay, Kate C. Fickas and MacKenzie O. Friedrichs
Remote Sens. 2024, 16(7), 1297; https://doi.org/10.3390/rs16071297 - 6 Apr 2024
Viewed by 1456
Abstract
The operational Simplified Surface Energy Balance (SSEBop) model has been utilized to generate gridded evapotranspiration data from Landsat images. These estimates are primarily driven by two sources of information: reference evapotranspiration and Landsat land surface temperature (LST) values. Hence, SSEBop is limited by [...] Read more.
The operational Simplified Surface Energy Balance (SSEBop) model has been utilized to generate gridded evapotranspiration data from Landsat images. These estimates are primarily driven by two sources of information: reference evapotranspiration and Landsat land surface temperature (LST) values. Hence, SSEBop is limited by the availability of Landsat data. Here, in this proof-of-concept paper, we utilize the Continuous Change Detection and Classification (CCDC) algorithm to generate synthetic Landsat data, which are then used as input for SSEBop to generate evapotranspiration estimates for six target areas in the continental United States, representing forests, shrublands, and irrigated agriculture. These synthetic land cover data are then used to generate the LST data required for SSEBop evapotranspiration estimates. The synthetic LST, evaporative fractions, and evapotranspiration data from CCDC closely mirror the phenological cycles in the observed Landsat data. Across the six sites, the median correlation in seasonal LST was 0.79, and the median correlation in seasonal evapotranspiration was 0.8. The median root mean squared error (RMSE) values were 2.82 °C for LST and 0.50 mm/day for actual evapotranspiration. CCDC predictions typically underestimate the average evapotranspiration by less than 1 mm/day. The average performance of the CCDC evaporative fractions, and corresponding evapotranspiration estimates, were much better than the initial LST estimates and, therefore, promising. Future work could include bias correction to improve CCDC’s ability to accurately reproduce synthetic Landsat data during the summer, allowing for more accurate evapotranspiration estimates, and determining the ability of SSEBop to predict regional evapotranspiration at seasonal timescales based on projected land cover change from CCDC. Full article
Show Figures

Figure 1

Figure 1
<p>The locations of the six target areas (white dots) in the western United States and their associated ecoregions. The Level I ecoregion data are available at ecologicalregions.info/htm/na_eco.htm (accessed on 23 February 2024).</p>
Full article ">Figure 2
<p>Red, green, and blue composite Landsat images of the six target areas. The Arizona target area (<b>a</b>) consists of irrigated croplands and desert. The California target area (<b>b</b>) consist mostly of irrigated almond plantations. Both the Colorado (<b>c</b>) and Oregon (<b>d</b>) target areas are largely forested. Then, both the Idaho (<b>e</b>) and New Mexico (<b>f</b>) target areas are shrubland.</p>
Full article ">Figure 3
<p>Target-area-wide averages of the Landsat (purple) and synthetic (cyan) land surface temperature (LST) estimates for 2019. The yellow difference line represents the Landsat data minus the synthetic data at each timestep. Note that the gaps in both the Landsat and synthetic data were the result of applying the same Landsat cloud mask to both datasets.</p>
Full article ">Figure 4
<p>Heat plots of synthetic land surface temperature (LST), derived from Continuous Change Detection and Classification (CCDC) and Landsat LST estimations for 2019, showing a per-pixel comparison for each target area. The blue 1:1 line represents the hypothetical perfect fit between observed LST and synthetic LST. Note the logarithmic scale, with purple indicating high pixel counts and orange indicating low pixel counts. Linear regression results are indicated by the R<sup>2</sup> values and are significant at <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Target-area-wide averages of the Landsat (purple) and synthetic (cyan) evaporative fraction (<span class="html-italic">ETf</span>) estimates for 2019. The yellow difference line represents the Landsat data minus the synthetic data at each timestep. Note that the gaps in both the Landsat and synthetic data were the result of applying the same Landsat cloud mask to both datasets.</p>
Full article ">Figure 6
<p>Heat plots of synthetic evaporative fraction (<span class="html-italic">ETf</span>), derived from Continuous Change Detection and Classification (CCDC) and Landsat <span class="html-italic">ETf</span> estimations for 2019, showing a per-pixel comparison for each target area. The blue 1:1 line represents the hypothetical perfect fit between observed <span class="html-italic">ETf</span> and synthetic <span class="html-italic">ETf</span>. Note the logarithmic scale, with green indicating high pixel counts and brown indicating low pixel counts. Linear regression results are indicated by the R<sup>2</sup> values and are significant at <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Side-by-side comparisons of the synthetic actual evapotranspiration (<span class="html-italic">ETa</span>) data (<b>left</b>) derived from the Continuous Change Detection and Classification (CCDC) and the observed <span class="html-italic">ETa</span> from Landsat (<b>right</b>). The Arizona (<b>top</b>) data correspond to a Landsat overpass on 5 July 2019, and New Mexico (<b>bottom</b>) data are from an overpass on 30 June 2019. The location of each target area is indicated by the plus sign in the bottom left corner.</p>
Full article ">Figure 8
<p>Target-area-wide averages of the Landsat (purple) and synthetic (cyan) actual evapotranspiration (<span class="html-italic">ETa</span>) estimates for 2019. The yellow difference line represents the Landsat data minus the synthetic data at each timestep. Note that the gaps in both the Landsat and synthetic data were the result of applying the same Landsat cloud mask to both datasets.</p>
Full article ">Figure 9
<p>Heat plots of synthetic actual evapotranspiration (<span class="html-italic">ETa</span>), derived from Continuous Change Detection and Classification (CCDC), and Landsat <span class="html-italic">ETa</span> estimations for 2019, showing a per-pixel comparison for each target area. The blue 1:1 line represents the hypothetical perfect fit between observed <span class="html-italic">ETa</span> and synthetic <span class="html-italic">ETa</span>. Note the logarithmic scale, with green indicating high pixel counts and purple indicating low pixel counts. Linear regression results are indicated by the R<sup>2</sup> values and are significant at <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 10
<p>Actual evapotranspiration (<span class="html-italic">ETa</span>) bias map showing the synthetic <span class="html-italic">ETa</span> estimates minus the Landsat <span class="html-italic">ETa</span> estimates for 21 July 2019, showing the large amount of spatial variation due to irrigation. The synthetic data underestimated <span class="html-italic">ETa</span> for fields that had recently been irrigated, and high <span class="html-italic">ETa</span> in the Landsat data (blue) and overestimated <span class="html-italic">ETa</span> (red) in agricultural fields that had not been irrigated recently.</p>
Full article ">
19 pages, 9215 KiB  
Article
Changes in Vegetation NDVI and Its Response to Climate Change and Human Activities in the Ferghana Basin from 1982 to 2015
by Heli Zhang, Lu Li, Xiaoen Zhao, Feng Chen, Jiachang Wei, Zhimin Feng, Tiyuan Hou, Youping Chen, Weipeng Yue, Huaming Shang, Shijie Wang and Mao Hu
Remote Sens. 2024, 16(7), 1296; https://doi.org/10.3390/rs16071296 - 6 Apr 2024
Cited by 7 | Viewed by 1303
Abstract
Exploring the evolution of vegetation cover and its drivers in the Ferghana Basin helps to understand the current ecological status of the Ferghana Basin and to analyze the vegetation changes and drivers, with a view to providing a scientific basis for regional ecological [...] Read more.
Exploring the evolution of vegetation cover and its drivers in the Ferghana Basin helps to understand the current ecological status of the Ferghana Basin and to analyze the vegetation changes and drivers, with a view to providing a scientific basis for regional ecological and environmental management and planning. Based on GIMMS NDVI3g and meteorological data, the spatial and temporal evolution characteristics of NDVI were analyzed from multiple perspectives with the help of linear trend and Mann–Kendall (MK) test methods using arcgis and the R language spatial analysis module, combined with partial correlation coefficients and residual analysis methods to analyze the impacts of climate change and human activities on the regional vegetation cover of the Ferghana Basin from 1982 to 2015. NDVI driving forces. The results showed the following: (1) The growing season of vegetation NDVI in the Ferghana Basin showed an increasing trend in the 34-year period, with an increase rate of 0.0044/10a, and the spatial distribution was significantly different, which was high in the central part of the country and low in the northern and southern parts of the country. (2) Temperature and precipitation simultaneously co-influenced the vegetation NDVI growth season, with most of the temperature and precipitation contributing in the spring, most of the temperature in the summer being negatively phased and the precipitation positively correlated, and most of the temperature and precipitation in the fall inhibiting vegetation NDVI growth. (3) The combined effect of climate change and human activities was the main reason for the overall rapid increase and great spatial variations in vegetation NDVI in China, and the spatial distribution of drivers, namely human activities and climate change, contributed 44.6% to vegetation NDVI in the growing season. The contribution of climate change and human activities to vegetation NDVI in the Ferghana Basin was 62.32% and 93.29%, respectively. The study suggests that more attention should be paid to the role of human activities and climate change in vegetation restoration to inform ecosystem management and green development. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area in the Ferghana Basin.</p>
Full article ">Figure 2
<p>Thirty-meter land cover types in the Ferghana Basin, 2010, 10—Rainfed cropland, 11—Herbaceous cover, 20—Irrigated cropland, 61—Open deciduous broadleaved forest, 62—Closed deciduous broadleaved forest, 71—Open evergreen needle-leaved forest, 72—Closed evergreen needle-leaved forest, 81—Open deciduous needle-leaved forest, 82—Closed deciduous needle-leaved forest, 120—Shrubland, 122—Deciduous shrubland, 130—Grassland, 140—Lichens and mosses, 150—Sparse vegetation, 180—Wetlands, 190—Impervious surfaces, 200—Bare areas, 201—Consolidated bare areas, 202—Unconsolidated bare areas, 210—Water body, 220—Permanent ice and snow.</p>
Full article ">Figure 3
<p>Interannual change of NDVI during developing season in the Ferghana Basin from 1982 to 2015.</p>
Full article ">Figure 4
<p>Spatial distribution of multi-year mean NDVI in vegetation growing season from 1982 to 2015.</p>
Full article ">Figure 5
<p>Significant distribution of NDVI changes in vegetation growing season from 1982 to 2015. (<b>A</b>): NDVI trend, (<b>B</b>): NDVI Trend Significance.</p>
Full article ">Figure 6
<p>Coefficient of variation of NDVI in the study area, 1982–2015.</p>
Full article ">Figure 7
<p>Interannual and growing season variations in climate factors in the study area, 1982−2015. (<b>A</b>): April−October factors; (<b>B</b>): Annual factors.</p>
Full article ">Figure 8
<p>Spatial distribution of biased correlations between NDVI and climate factors for growing season vegetation from 1982−2015. (<b>A</b>): Growing season NDVI is biased with precipitation; (<b>B</b>): Growing season NDVI is biased with temperature.</p>
Full article ">Figure 9
<p>(<b>A</b>–<b>C</b>): Spring, summer, and autumn vegetation NDVI bias correlation with precipitation; (<b>D</b>–<b>F</b>): Spring, summer, and autumn vegetation NDVI bias correlation with temperature.</p>
Full article ">Figure 10
<p>Spatial distribution of the impacts of climatic change and human activities on vegetation restoration in Ferghana Basin during 1982–2015. (<b>A</b>): Climate change; (<b>B</b>): Human activities.</p>
Full article ">Figure 11
<p>Spatial distribution of driving factors of vegetation cover change in the Ferghana Basin from 1982 to 2015 (CC and HA refer to climate change and human activities, respectively), ↑ represents an increase, ↓ represents a decrease.</p>
Full article ">Figure 12
<p>Spatial distribution of the contribution rate of climate change and human activities to vegetation cover change in the Ferghana Basin from 1982 to 2015. (<b>A</b>): Climate change; (<b>B</b>): Human activities.</p>
Full article ">
21 pages, 2358 KiB  
Article
Three-Dimensional Human Pose Estimation from Micro-Doppler Signature Based on SISO UWB Radar
by Xiaolong Zhou, Tian Jin, Yongpeng Dai, Yongping Song and Kemeng Li
Remote Sens. 2024, 16(7), 1295; https://doi.org/10.3390/rs16071295 - 6 Apr 2024
Cited by 2 | Viewed by 1303
Abstract
In this paper, we propose an innovative approach for transforming 2D human pose estimation into 3D models using Single Input–Single Output (SISO) Ultra-Wideband (UWB) radar technology. This method addresses the significant challenge of reconstructing 3D human poses from 1D radar signals, a task [...] Read more.
In this paper, we propose an innovative approach for transforming 2D human pose estimation into 3D models using Single Input–Single Output (SISO) Ultra-Wideband (UWB) radar technology. This method addresses the significant challenge of reconstructing 3D human poses from 1D radar signals, a task traditionally hindered by low spatial resolution and complex inverse problems. The difficulty is further exacerbated by the ambiguity in 3D pose reconstruction, as multiple 3D poses may correspond to similar 2D projections. Our solution, termed the Radar PoseLifter network, leverages the micro-Doppler signatures inherent in 1D radar echoes to effectively convert 2D pose information into 3D structures. The network is specifically designed to handle the long-range dependencies present in sequences of 2D poses. It employs a fully convolutional architecture, enhanced with a dilated temporal convolutions network, for efficient data processing. We rigorously evaluated the Radar PoseLifter network using the HPSUR dataset, which includes a diverse range of human movements. This dataset comprises data from five individuals with varying physical characteristics, performing a variety of actions. Our experimental results demonstrate the method’s robustness and accuracy in estimating complex human poses, highlighting its effectiveness. This research contributes significantly to the advancement of human motion capture using radar technology. It presents a viable solution for applications where precision and reliability in motion capture are paramount. The study not only enhances the understanding of 3D pose estimation from radar data but also opens new avenues for practical applications in various fields. Full article
(This article belongs to the Special Issue Advances in Remote Sensing, Radar Techniques, and Their Applications)
Show Figures

Figure 1

Figure 1
<p>Configuration of SISO UWB bistatic radar for human pose estimation showing the relative positioning of the transmit and receive antennas, along with the azimuth and elevation angles to the moving human target.</p>
Full article ">Figure 2
<p>The geometric relationship between human motion model and radar.</p>
Full article ">Figure 3
<p>Schematic representation of the proportional human skeletal structure.</p>
Full article ">Figure 4
<p>Simulated experimental scene from MOCAP data.</p>
Full article ">Figure 5
<p>The micro-Doppler spectrum of human motion.</p>
Full article ">Figure 6
<p>The micro-Doppler spectrum of individual human body parts for movement analysis.</p>
Full article ">Figure 7
<p>Radar data preprocessing chain chart.</p>
Full article ">Figure 8
<p>The micro-Doppler spectrum of four different human activities of the HPSUR dataset.</p>
Full article ">Figure 9
<p>Schematic overview of the methodological framework for 3D human pose estimation.</p>
Full article ">Figure 10
<p>Illustration of data collection scenarios of HPSUR dataset.</p>
Full article ">Figure 11
<p>Comparative visualization of 3D human pose reconstruction from radar-derived 2D keypoints.</p>
Full article ">
24 pages, 3929 KiB  
Article
Sentinel-1-Based Soil Freeze–Thaw Detection in Agro-Forested Areas: A Case Study in Southern Québec, Canada
by Shahabeddin Taghipourjavi, Christophe Kinnard and Alexandre Roy
Remote Sens. 2024, 16(7), 1294; https://doi.org/10.3390/rs16071294 - 6 Apr 2024
Cited by 1 | Viewed by 1314
Abstract
Nearly 50 million km2 of global land experiences seasonal transitions from predominantly frozen to thawed conditions, significantly impacting various ecosystems and hydrologic processes. In this study, we assessed the capability to retrieve surface freeze–thaw (FT) conditions using Sentinel-1 synthetic aperture radar (SAR) [...] Read more.
Nearly 50 million km2 of global land experiences seasonal transitions from predominantly frozen to thawed conditions, significantly impacting various ecosystems and hydrologic processes. In this study, we assessed the capability to retrieve surface freeze–thaw (FT) conditions using Sentinel-1 synthetic aperture radar (SAR) data time series at two agro-forested study sites, St-Marthe and St-Maurice, in southern Québec, Canada. In total, 18 plots were instrumented to monitor soil temperature and derive soil freezing probabilities at 2 and 10 cm depths during 2020–21 and 2021–22. Three change detection algorithms were tested: backscatter differences (∆σ) derived from thawed reference (Delta), the freeze–thaw index (FTI), and a newly developed exponential freeze–thaw algorithm (EFTA). Various probabilistic mixed models were compared to identify the model and predictor variables that best predicted soil freezing probability. VH polarization backscatter signals processed with the EFTA and used as predictors in a logistic model led to improved predictions of soil freezing probability at 2 cm (Pseudo-R2 = 0.54) compared to other approaches. The EFTA could effectively address the limitations of the Delta algorithm caused by backscatter fluctuations in the shoulder seasons, resulting in more precise estimates of FT events. Furthermore, the inclusion of crop types as plot-level effects within the probabilistic model also slightly improved the soil freezing probability prediction at each monitored plot, with marginal and conditional R2 values of 0.59 and 0.61, respectively. The model accurately classified observed binary ‘frozen’ or ‘thawed’ states with 85.2% accuracy. Strong cross-level interactions were also observed between crop types and the EFTA derived from VH backscatter, indicating that crop type modulated the backscatter response to soil freezing. This study represents the first application of the EFTA and a probabilistic approach to detect frozen soil conditions in agro-forested areas in southern Quebec, Canada. Full article
(This article belongs to the Special Issue Radar Remote Sensing for Monitoring Agricultural Management)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overview of the study sites with land cover map in 2020 (<a href="https://open.canada.ca/data/en/dataset/ee1580ab-a23d-4f86-a09b-79763677eb47" target="_blank">https://open.canada.ca/data/en/dataset/ee1580ab-a23d-4f86-a09b-79763677eb47</a>, accessed on 1 April 2024). (<b>a</b>) Geographical depiction of the study area located in Canada. (<b>b</b>) The geographical location of study sites in the agro-forested areas in southern Québec. (<b>c</b>) Study plot locations in St-Maurice (six in farmlands and two in forest). (<b>d</b>) Study plot locations in St-Marthe (eight in agricultural lands and two in forest). The map was created using ArcGIS version 10.3 software (Esri, Redlands, CA, USA).</p>
Full article ">Figure 2
<p>Field observations of crop residues at both sites during the study periods 2020–21 and 2021–22. (<b>a</b>,<b>b</b>) Grass presence in St-Marthe’s C plot over the study period. (<b>c</b>) Snow-covered grass in St-Marthe’s E plot. (<b>d</b>) Snow-free grass in St-Marthe’s F plot. (<b>e</b>) Plowed soils in St-Marthe’s G plot. (<b>f</b>) Plowed soils in St-Marthe’s H plot. (<b>g</b>) Corn stalks residues and scattered debris in St-Marthe’s I plot. (<b>h</b>) Corn stalks residues and scattered debris in St-Marthe’s J plot. (<b>i</b>,<b>j</b>) Fields with bare lands and corn stalks scattered throughout St-Maurice’s C, D, and E plots.</p>
Full article ">Figure 3
<p>Spatial and temporal variations in freezing probability at 2 cm depth along with corresponding S1 overpasses for all study plots. (<b>a</b>,<b>b</b>) St-Marthe for 2020–21 and 2021–22. (<b>c</b>,<b>d</b>) St-Maurice for 2020–21 and 2021–22. A and B plots in each site located in forest (green rectangles). For St-Marthe’s J and I and St-Maurice’s F, G, and H plots, the initiation of soil temperature monitoring started later, resulting in an absence of freezing probability values.</p>
Full article ">Figure 4
<p>Comparison of observed 2 cm soil freezing probabilities (dark blue, right Y-axis) with that predicted by the Delta method (top row, in red) and EFTA (bottom row, in red). The S1 corrected VH polarization backscatter signal is also displayed in green. On the left Y-axis, values less than 0 correspond to corrected VH backscatters (green curve), while values greater than 0 correspond to the predicted freezing probability (red curves). (<b>a</b>,<b>b</b>) St-Maurice’s H plot in 2020–21. (<b>c</b>,<b>d</b>) St-Maurice’s H plot in 2021–22. (<b>e</b>,<b>f</b>) St-Marthe’s H plot in 2020–21. (<b>g</b>,<b>h</b>) St-Marthe’s H plot in 2021–22. (<b>i</b>,<b>j</b>) St-Marthe’s I plot in 2020–21. (<b>k</b>,<b>l</b>) St-Marthe’s I plot in 2021–22. The R<sup>2</sup> values, derived from the correlation between the soil freezing probability at 2 cm depth and VH<sub>Delta</sub> (top plots)/and VH<sub>EFTA</sub> (bottom plots). A dashed line represents the overall trend between VH<sub>Delta</sub> and the date (top plots) or VH<sub>EFTA</sub> and the date (bottom plots).</p>
Full article ">Figure 5
<p>Local fitted logistic models for observed freezing probability (2 cm) against VH<sub>EFTA</sub> in St-Marthe (top) and St-Maurice (bottom). (<b>a</b>) St-Marthe’s H agricultural plot. (<b>b</b>) St-Marthe’s I agricultural plot. (<b>c</b>) St-Marthe’s B forest plot. (<b>d</b>) St-Maurice’s F agricultural plot. (<b>e</b>) St-Maurice’s H agricultural plot. (<b>f</b>) St-Maurice’s A forest plot. The top and bottom Pseudo-R<sup>2</sup> (Pse-R<sup>2</sup>) values of each plot are shown for 2020–21 and 2021–22.</p>
Full article ">Figure 6
<p>Cross-over interactions of different crop types on the relationship between VH<sub>EFTA</sub> and the predicted soil freezing probability.</p>
Full article ">Figure 7
<p>Spatial cross-validation results of the Crop-Mixed soil freezing probability model with 28 single folds, showing 14 plots evaluated across two study years: 2020–21 (gray) and 2021–22 (white). The figure displays the Brier score, MAE, and R<sup>2</sup> for each fold when individually excluded from calibration and used for validation. The statistical metrics averages for the plots across the study years 2020–21 and 2021–22 exhibited the following respective values: For 2020–21: R<sup>2</sup> = 0.55, Brier Score = 0.19, and MAE = 0.18; and for 2021–22: R<sup>2</sup> = 0.60, Brier Score = 0.18, and MAE = 0.17.</p>
Full article ">Figure 8
<p>Comparison of the soil freezing probability predicted by the Crop-Mixed model vs. observed values. (<b>a</b>) Scatter plot of predicted vs. observed data; (<b>b</b>) frequency histogram showing the accuracy (true positive/negative) and misclassification (false positive/negative) for the different crop types. In the left plot, the red line depicts the best fit for the plotted points, while the gray line illustrates the hypothetical scenario where predicted values perfectly align with observed ones. A smaller gap between these lines signifies the higher model's performance.</p>
Full article ">
17 pages, 4146 KiB  
Article
CDEST: Class Distinguishability-Enhanced Self-Training Method for Adopting Pre-Trained Models to Downstream Remote Sensing Image Semantic Segmentation
by Ming Zhang, Xin Gu, Ji Qi, Zhenshi Zhang, Hemeng Yang, Jun Xu, Chengli Peng and Haifeng Li
Remote Sens. 2024, 16(7), 1293; https://doi.org/10.3390/rs16071293 - 6 Apr 2024
Viewed by 920
Abstract
The self-supervised learning (SSL) technique, driven by massive unlabeled data, is expected to be a promising solution for semantic segmentation of remote sensing images (RSIs) with limited labeled data, revolutionizing transfer learning. Traditional ‘local-to-local’ transfer from small, local datasets to another target dataset [...] Read more.
The self-supervised learning (SSL) technique, driven by massive unlabeled data, is expected to be a promising solution for semantic segmentation of remote sensing images (RSIs) with limited labeled data, revolutionizing transfer learning. Traditional ‘local-to-local’ transfer from small, local datasets to another target dataset plays an ever-shrinking role due to RSIs’ diverse distribution shifts. Instead, SSL promotes a ‘global-to-local’ transfer paradigm, in which generalized models pre-trained on arbitrarily large unlabeled datasets are fine-tuned to the target dataset to overcome data distribution shifts. However, the SSL pre-trained models may contain both useful and useless features for the downstream semantic segmentation task, due to the gap between the SSL tasks and the downstream task. To adapt such pre-trained models to semantic segmentation tasks, traditional supervised fine-tuning methods that use only a small number of labeled samples may drop out useful features due to overfitting. The main reason behind this is that supervised fine-tuning aims to map a few training samples from the high-dimensional, sparse image space to the low-dimensional, compact semantic space defined by the downstream labels, resulting in a degradation of the distinguishability. To address the above issues, we propose a class distinguishability-enhanced self-training (CDEST) method to support global-to-local transfer. First, the self-training module in CDEST introduces a semi-supervised learning mechanism to fully utilize the large amount of unlabeled data in the downstream task to increase the size and diversity of the training data, thus alleviating the problem of biased overfitting of the model. Second, the supervised and semi-supervised contrastive learning modules of CDEST can explicitly enhance the class distinguishability of features, helping to preserve the useful features learned from pre-training while adapting to downstream tasks. We evaluate the proposed CDEST method on four RSI semantic segmentation datasets, and our method achieves optimal experimental results on all four datasets compared to supervised fine-tuning as well as three semi-supervised fine-tuning methods. Full article
Show Figures

Figure 1

Figure 1
<p>Overview framework of the class distinguishability-enhanced self-training method.</p>
Full article ">Figure 2
<p>Scheme of supervised contrastive learning module for semantic segmentation.</p>
Full article ">Figure 3
<p>The self-training module and semi-supervised contrastive learning module.</p>
Full article ">Figure 4
<p>Comparison of visual results from four RSI semantic segmentation tasks. Where the two RSIs in (<b>a</b>) are from the Potsdam dataset, the two RSIs in (<b>b</b>) are from the DGLCC dataset, the two RSIs in (<b>c</b>) are from the Hubei dataset, and the two RSIs in (<b>d</b>) are from the Xiangtan dataset. * Fine-tuning methods that use the multi-round iterative self-training strategy continue to train until accuracy degrades.</p>
Full article ">Figure 5
<p>Visualization of randomly sampled pixel features’ t-SNE for each category: (<b>a</b>) Without <math display="inline"><semantics> <msubsup> <mi mathvariant="bold-script">L</mi> <mrow> <mi mathvariant="bold-italic">con</mi> </mrow> <mi mathvariant="bold-italic">lbl</mi> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi mathvariant="bold-script">L</mi> <mrow> <mi mathvariant="bold-italic">con</mi> </mrow> <mi mathvariant="bold-italic">un</mi> </msubsup> </semantics></math>, (<b>b</b>) Complete method.</p>
Full article ">Figure 6
<p>Comparative results on Potsdam and Xiangtang datasets using different numbers of unlabeled data for fine-tuning.</p>
Full article ">
20 pages, 51888 KiB  
Article
Introducing the Azimuth Cutoff as an Independent Measure for Characterizing Sea-State Dynamics in SAR Altimetry
by Ourania Altiparmaki, Samira Amraoui, Marcel Kleinherenbrink, Thomas Moreau, Claire Maraldi, Pieter N. A. M. Visser and Marc Naeije
Remote Sens. 2024, 16(7), 1292; https://doi.org/10.3390/rs16071292 - 6 Apr 2024
Cited by 2 | Viewed by 996
Abstract
This study presents the first azimuth cutoff analysis in Synthetic Aperture Radar (SAR) altimetry, aiming to assess its applicability in characterizing sea-state dynamics. In SAR imaging, the azimuth cutoff serves as a proxy for the shortest waves, in terms of wavelength, that can [...] Read more.
This study presents the first azimuth cutoff analysis in Synthetic Aperture Radar (SAR) altimetry, aiming to assess its applicability in characterizing sea-state dynamics. In SAR imaging, the azimuth cutoff serves as a proxy for the shortest waves, in terms of wavelength, that can be detected by the satellite under certain wind and wave conditions. The magnitude of this parameter is closely related to the wave orbital velocity variance, a key parameter for characterizing wind-wave systems. We exploit wave modulations exhibited in the tail of fully-focused SAR waveforms and extract the azimuth cutoff from the radar signal through the analysis of its along-track autocorrelation function. We showcase the capability of Sentinel-6A in deriving these two parameters based on analyses in the spatial and wavenumber domains, accompanied by a discussion of the limitations. We use Level-1A high-resolution Sentinel-6A data from one repeat cycle (10 days) globally to verify our findings against wave modeled data. In the spatial domain analysis, the estimation of azimuth cutoff involves fitting a Gaussian function to the along-track autocorrelation function. Results reveal pronounced dependencies on wind speed and significant wave height, factors primarily determining the magnitude of the velocity variance. In extreme sea states, the parameters are underestimated by the altimeter, while in relatively calm sea states and in the presence of swells, a substantial overestimation trend is observed. We introduce an alternative approach to extract the azimuth cutoff by identifying the fall-off wavenumber in the wavenumber domain. Results indicate effective mitigation of swell-induced errors, with some additional sensitivity to extreme sea states compared to the spatial domain approach. Full article
(This article belongs to the Special Issue Advances in Satellite Altimetry II)
Show Figures

Figure 1

Figure 1
<p>Examples of Sentinel-6A fully-focused SAR waveform-tail radargrams. Wind and wave conditions, given in the title of each panel, obtained from collocated ERA5 products.</p>
Full article ">Figure 2
<p>Interpolated significant wave height (<b>left column</b>) and mean zero-up crossing period (<b>right column</b>) parameters of MFWAM (<b>top row</b>) and ERA5 (<b>bottom row</b>) to Sentinel-6A tracks.</p>
Full article ">Figure 3
<p>Ratio of the <math display="inline"><semantics> <msubsup> <mi>σ</mi> <mrow> <msub> <mi>υ</mi> <mrow> <mi>w</mi> <mi>w</mi> <mo>,</mo> <mi>h</mi> <mi>f</mi> </mrow> </msub> </mrow> <mn>2</mn> </msubsup> </semantics></math> wave orbital velocity variance of high-frequency waves (&gt;0.58 Hz) to the <math display="inline"><semantics> <msubsup> <mi>σ</mi> <mrow> <msub> <mi>υ</mi> <mrow> <mi>w</mi> <mi>w</mi> </mrow> </msub> </mrow> <mn>2</mn> </msubsup> </semantics></math> total wave orbital velocity variance versus wind speed. The colors represent the different fetches.</p>
Full article ">Figure 4
<p>Example of the along-track autocorrelation function, depicted in gray color, of a Sentinel-6A (S6A) fully-focused SAR waveform-tail radargram acquired for a scene characterized by moderate wind and wave conditions. The blue dashed line represents the Gaussian function. The yellow dash-dotted lines represent the Sentinel-6A azimuth cutoff estimate. Sea-state conditions are obtained from ERA5 products.</p>
Full article ">Figure 5
<p>Scatter plots of the azimuth cutoff estimated by Sentinel-6A in comparison to ERA5 (<b>top panels</b>) and MFWAM (<b>bottom panels</b>). In the left and right panels, color gradations correspond to wind speed and significant wave height, respectively. The square markers linked with solid gray lines depict the average value of points grouped every 50 m considering wave model-derived intervals. The gray dashed lines represent the scenario where the azimuth cutoff of the satellite and wave model would align perfectly.</p>
Full article ">Figure 6
<p>Examples of Sentinel-6A fully-focused SAR waveform-tail radargrams (<b>left panels</b>) and their along-track autocorrelation functions (<b>right panels</b>), illustrated in gray color. The blue dashed line represents the Gaussian function. The yellow, pink, and brown dash-dotted lines represent the Sentinel-6A, ERA5 and MFWAM azimuth cutoff estimates, respectively. Sea state conditions are obtained from ERA5 products.</p>
Full article ">Figure 7
<p>Examples of along-track autocorrelation functions, depicted in gray color, concerning SAR waveform-tail radargrams dominated by swells traveling in the cross-track direction (<b>top left</b>), at angle (<b>top right</b> and <b>bottom left</b>) and in the along-track direction (<b>bottom right</b>). The blue dashed line represents the Gaussian function. The yellow, pink, and brown dash-dotted lines represent the Sentinel-6A, ERA5 and MFWAM azimuth cutoff estimates, respectively. Sea-state conditions are obtained from ERA5 products.</p>
Full article ">Figure 8
<p>Example of the spectral autocorrelation function (gray solid line). The yellow circular marker represents the fall-off wavenumber identified as the intersection point between the gray dotted line, depicting the threshold line equal to five times the median spectral along-track autocorrelation function, and the blue solid line representing the seventh-order polynomial model. The pink and brown dash-dotted lines represent the fall-off wavenumber estimates from ERA5 and MFWAM, respectively. The yellow dash-dotted represents the Sentinel-6A fall-off wavenumber obtained from the spatial domain-derived azimuth cutoff estimate. Sea-state conditions are obtained from ERA5 products.</p>
Full article ">Figure 9
<p>Global maps of the azimuth cutoff as derived by ERA5 (<b>top left</b>), MFWAM (<b>top right</b>), Sentinel-6A from the spatial domain (SD) analysis (<b>bottom left</b>), and Sentinel-6A from the wavenumber domain (or Fourier Domain—FD) analysis (<b>bottom right</b>). Each map is accompanied by a histogram showing the distribution of the estimates on a logarithmic scale.</p>
Full article ">Figure 10
<p>Global maps of the velocity variance as derived by ERA5 (<b>top left</b>), MFWAM (<b>top right</b>), Sentinel-6A from the spatial domain (SD) analysis (<b>bottom left</b>), and Sentinel-6A from the wavenumber domain (or Fourier Domain—FD) analysis (<b>bottom right</b>). Each map is accompanied by a histogram showing the distribution of the estimates on a logarithmic scale.</p>
Full article ">Figure 11
<p>Global maps of the peak wave period and wind speed as obtained from ERA5 (<b>top panels</b>) and the velocity variance differences between ERA5 and Sentinel-6A in the spatial domain (SD) (<b>mid left</b>) and wavenumber domain (or Fourier Domain—FD) (<b>mid right</b>), MFWAM and Sentinel-6A in the spatial (<b>bottom left</b>) and wavenumber (<b>bottom right</b>) domains. Histograms of velocity variance differences accompany the middle and bottom panels.</p>
Full article ">
19 pages, 14368 KiB  
Communication
Sub-Hourly Variations of Wind Shear in the Mesosphere-Lower Thermosphere as Observed by the China Meteor Radar Chain
by Chi Long, Tao Yu, Jian Zhang, Xiangxiang Yan, Na Yang, Jin Wang, Chunliang Xia, Yu Liang and Hailun Ye
Remote Sens. 2024, 16(7), 1291; https://doi.org/10.3390/rs16071291 - 6 Apr 2024
Viewed by 675
Abstract
Wind shear has important implications for Kelvin–Helmholtz instability (KHI) and gravity waves (GWs) in the mesosphere–lower thermosphere (MLT) region where its momentum transport process is dominated by short-period (<1 h) GWs. However, the sub-hourly variation in wind shear is still not well quantified. [...] Read more.
Wind shear has important implications for Kelvin–Helmholtz instability (KHI) and gravity waves (GWs) in the mesosphere–lower thermosphere (MLT) region where its momentum transport process is dominated by short-period (<1 h) GWs. However, the sub-hourly variation in wind shear is still not well quantified. This study aims to improve current understanding of vertical wind shear by analyzing multi-year meteor radar measurements at the Mohe (MH, 53.5°N, 122.3°E), Beijing (BJ, 40.3°N, 116.2°E), Wuhan (WH, 30.5°N, 114.6°E), and Fuke (FK, 19.5°N, 109.1°E) stations in China. The wind field is estimated by a new algorithm, e.g., the damped least squares fitting. Taking the wind shear estimated by normal products as a criterion, the shear produced by the new algorithm has more statistical convergence as compared to the traditional algorithm, e.g., the least squares fitting. Therefore, we argue that the 10 min DLSA wind probably produces a more reasonable vertical shear. Both intensive wind shears and GW kinetic energy can be simultaneously captured during the 0600–1600 UTs of May at MH and during the 1300–2400 UTs of March at FK, possibly implying that the up-propagation of GWs could contribute to the production of large wind shears. The sub-hourly variation in wind shears is potentially valuable for understanding the interrelationship between shear (or KHI) and GWs. Full article
Show Figures

Figure 1

Figure 1
<p>The locations of four meteor radar stations used in this study. The station names are Mohe (MH), Beijing (BJ), Wuhan (WH), and Fuke (FK), respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>d</b>) Daily total meteor echoes detected by the MH, BJ, WH, and FK radars. The red vertical lines mark the period from 20 to 22 September 2018. (<b>e</b>) The height distribution of meteor echoes is measured in 1 km bins on 20 September 2018 at MH, with a Gaussian fitting represented by the red curve. The meteor peak height and its standard deviation are marked. Similarly, (<b>f</b>) indicates the result from 21 September 2018, (<b>g</b>) indicates the result from 22 September 2018, and (<b>h</b>) indicates the hourly distribution of meteor echoes from 20 September 2018 at 88 and 90 km at MH, respectively. Similarly, (<b>i</b>) indicates the result from 21 September 2018 and (<b>j</b>) indicates the result on 22 September 2018.</p>
Full article ">Figure 3
<p>An example showing the variations in 60 min and 10 min resolution zonal winds obtained by different algorithms during 20 to 22 September 2018 at the four stations. (<b>a</b>–<b>c</b>) indicate the 60 min LSA wind, 10 min LSA wind, and 10 min DLSA wind at MH, respectively. Similarly, (<b>d</b>–<b>f</b>) indicate the results at BJ, (<b>g</b>–<b>i</b>) indicate the results at WH, and (<b>j</b>–<b>l</b>) indicate the results at FK.</p>
Full article ">Figure 4
<p>Similar to <a href="#remotesensing-16-01291-f003" class="html-fig">Figure 3</a>, indicates the results in meridional winds.</p>
Full article ">Figure 5
<p>The comparison in estimated error for wind velocity between LSA and DLSA at different heights during 20 to 22 September 2018 at the MH station. The red (blue) line denotes a linear fit to the red (blue) dots, with the slope <math display="inline"><semantics> <mrow> <mo>±</mo> </mrow> </semantics></math> the uncertainty in the slope and the y-intercept <math display="inline"><semantics> <mrow> <mo>±</mo> </mrow> </semantics></math> the uncertainty in the y-intercept indicated correspondingly.</p>
Full article ">Figure 6
<p>Statistical results of estimated error for retrieved wind velocity between LSA and DLSA during the years 2013–2022 at MH. The box and whisker plots of estimated error for wind velocity over eight meteor count units (i.e., 5–10, 10–15, 15–20, 20–25, 25–30, 30–35, 35–40, and 40–45) over four heights are displayed in (<b>a</b>–<b>d</b>). The standard error in a mean is marked next to the box and whisker plot. In total, 100 times the standard error is shown by each distribution. One-way analysis of variance: 86 km, LSA P = 0 and DLSA P = 0; 88 km, LSA P = 0 and DLSA P = 0. 90 km, LSA P = 0 and DLSA P = 0; 92 km, LSA P = 0, and DLSA P = 0.</p>
Full article ">Figure 7
<p>Similar to <a href="#remotesensing-16-01291-f006" class="html-fig">Figure 6</a>, this figure indicates the results from BJ, WH, and FK. The standard error in a mean is marked next to the box and whisker plots. In total, 100 times the standard error is shown by each distribution. One-way analysis of variance: 86 km at BJ, LSA P = 0 and DLSA P = 0; 86 km at WH, LSA P = 0 and DLSA P = 0; 86 km at FK, LSA P = 0 and DLSA P = 0; 88 km at BJ, LSA P = 0 and DLSA P = 0. 88 km at WH, LSA P = 0 and DLSA P = 0. 88 km at FK, LSA P = 0 and DLSA P= 0. 90 km at BJ, LSA P = 0 and DLSA P = 0; 90 km at WH, LSA P = 0 and DLSA P = 0; 90 km at FK, LSA P = 0 and DLSA P = 0; 92 km at BJ, LSA P = 0 and DLSA P = 0. 92 km at WH, LSA P = 0 and DLSA P = 0. 92 km at FK, LSA P = 0, and DLSA P = 0.</p>
Full article ">Figure 8
<p>Comparison of LSA and DLSA as function of the correlation coefficient and cost function at different heights during 14 to 26 September 2018 at MH. The (blue and red) bars denote cost function (left Y-axis), and the (blue and red) lines denote correlation coefficient (right Y-axis).</p>
Full article ">Figure 9
<p>Statistical results of correlation coefficient and cost function between LSA and DLSA during the years 2013–2022 at MH are presented in (<b>a</b>) and (<b>b</b>), respectively. The box and whisker plots for the correlation coefficient (cost function) over four heights are displayed. The standard error in a mean is marked next to the box and whisker plots, while 100 times the standard error is shown by each correlation coefficient distribution. One-way analysis of variance: correlation coefficient, LSA P = 1.47 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>10</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 5.57 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>30</mn> </mrow> </msup> </mrow> </semantics></math>; cost function, LSA P = 1.2 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>96</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 3.67 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>100</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Similar to <a href="#remotesensing-16-01291-f009" class="html-fig">Figure 9</a>, this figure indicates the result from BJ, WH, and FK. The standard error in a mean is marked next to the box and whisker plots, while 100 times the standard error is shown by each correlation coefficient distribution. One-way analysis of variance: correlation coefficient at BJ, LSA P = 0.0131 and DLSA P = 1.99 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>5</mn> </mrow> </msup> </mrow> </semantics></math>; cost function at BJ, LSA P = 0.0021 and DLSA P = 5.29 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>. correlation coefficient at WH, LSA P = 9.8 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>41</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 1.92 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>58</mn> </mrow> </msup> </mrow> </semantics></math>; cost function at WH, LSA P = 2.79 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>11</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 9.12 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>14</mn> </mrow> </msup> </mrow> </semantics></math>. correlation coefficient at FK, LSA P = 7.40 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>10</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 1.78 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>13</mn> </mrow> </msup> </mrow> </semantics></math>; cost function at FK, LSA P = 7.19 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 1.59 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>10</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>The 10 min wind shears at MH calculated on (<b>a</b>) 20 September 2018 using LSA; (<b>b</b>) 20 September 2018 using DLSA; (<b>c</b>) 21 September 2018 using LSA; (<b>d</b>) 21 September 2018 using DLSA; (<b>e</b>) 22 September 2018 using LSA; and (<b>f</b>) 22 September 2018 using DLSA. The red dotted line stands for the 95th percentile of the 60 min wind shears during the years 2013–2022 at MH, and the effective ratio denotes the ratio between shears smaller than 21.3 m/s/km and all shears. The blue triangles (dots) represent the shears in LSA wind (DLSA wind) smaller than 21.3 m/s/km, and the red triangles (dots) represent the shears in LSA wind (DLSA wind) bigger than 21.3 m/s/km. The standard deviation is marked in the plot.</p>
Full article ">Figure 12
<p>The comparison of 10 min DLSA wind shears at different stations between 20 to 22 September 2018 and the three-month averaged value from September to November 2018. The blue line represents the shear value of each station on that day, and the red line represents the averaged shear value of each station.</p>
Full article ">Figure 13
<p>Statistical results of wind shear between the 10 min LSA wind and 10-min DLSA wind during the whole study period at four different stations. The box and whisker plots of wind shear over eight UTC units (i.e., 0–3, 3–6, 6–9, 9–12, 12–15, 15–18, 18–21, and 21–24) over four stations are displayed. The standard deviation is presented in the form of error bars. The mean shears of the 10 min LSA wind (or DLSA wind) are connected by a black fold line. The standard error in a mean is marked next to the box and whisker plots. One-way analysis of variance: at MH, LSA P = 7.81 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>52</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 3.52 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>69</mn> </mrow> </msup> </mrow> </semantics></math>; at BJ, LSA P = 2.79 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>11</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 1.23 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>39</mn> </mrow> </msup> </mrow> </semantics></math>; at WH, LSA P = 3.85 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>40</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 6.52 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>55</mn> </mrow> </msup> </mrow> </semantics></math>; at FK, LSA P = 3.79 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>21</mn> </mrow> </msup> </mrow> </semantics></math> and DLSA P = 4.17 <math display="inline"><semantics> <mrow> <mo>×</mo> <mo> </mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>29</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 14
<p>The distributions of the 10 min shear calculated by the DLSA at four radar sites as functions of UT and month. The estimated wind shear is averaged during the years 2013–2022 at MH (<b>a</b>), during the years 2018–2022 at BJ (<b>b</b>), during the years 2013–2022 at WH (<b>c</b>), and during the years 2018–2020 at FK (<b>d</b>). The blank spaces are due to a lack of meteor counts.</p>
Full article ">Figure 15
<p>The hour–month cross section of the GW kinetic energy calculated by the 10 min DLSA wind. The estimated energy is averaged during the years 2013–2022 at MH (<b>a</b>), during the years 2018–2022 at BJ (<b>b</b>), during the years 2013–2022 at WH (<b>c</b>), and during the years 2018–2020 at FK (<b>d</b>). The blank spaces are due to a lack of meteor counts.</p>
Full article ">
34 pages, 6717 KiB  
Article
Estimating Carbon Dioxide Emissions from Power Plant Water Vapor Plumes Using Satellite Imagery and Machine Learning
by Heather D. Couture, Madison Alvara, Jeremy Freeman, Aaron Davitt, Hannes Koenig, Ali Rouzbeh Kargar, Joseph O’Connor, Isabella Söldner-Rembold, André Ferreira, Jeyavinoth Jeyaratnam, Jordan Lewis, Colin McCormick, Tiffany Nakano, Charmaine Dalisay, Christy Lewis, Gabriela Volpato, Matthew Gray and Gavin McCormick
Remote Sens. 2024, 16(7), 1290; https://doi.org/10.3390/rs16071290 - 6 Apr 2024
Viewed by 2102
Abstract
Combustion power plants emit carbon dioxide (CO2), which is a major contributor to climate change. Direct emissions measurement is cost-prohibitive globally, while reporting varies in detail, latency, and granularity. To fill this gap and greatly increase the number of power plants [...] Read more.
Combustion power plants emit carbon dioxide (CO2), which is a major contributor to climate change. Direct emissions measurement is cost-prohibitive globally, while reporting varies in detail, latency, and granularity. To fill this gap and greatly increase the number of power plants worldwide with independent emissions monitoring, we developed and applied machine learning (ML) models using power plant water vapor plumes as proxy signals to estimate electric power generation and CO2 emissions using Landsat 8, Sentinel-2, and PlanetScope imagery. Our ML models estimated power plant activity on each image snapshot, then an aggregation model predicted plant utilization over a 30-day period. Lastly, emission factors specific to region, fuel, and plant technology were used to convert the estimated electricity generation into CO2 emissions. Models were trained with reported hourly electricity generation data in the US, Europe, and Australia and were validated with additional generation and emissions data from the US, Europe, Australia, Türkiye, and India. All results with sufficiently large sample sizes indicate that our models outperformed the baseline approaches. In validating our model results against available generation and emissions reported data, we calculated the root mean square error as 1.75 TWh (236 plants across 17 countries over 4 years) and 2.18 Mt CO2 (207 plants across 17 countries over 4 years), respectively. Ultimately, we applied our ML method to plants that constitute 32% of global power plant CO2 emissions, as estimated by Climate TRACE, averaged over the period 2015–2022. This dataset is the most comprehensive independent and free-of-cost global power plant point-source emissions monitoring system currently known to the authors and is made freely available to the public to support global emissions reduction. Full article
(This article belongs to the Special Issue Remote Sensing for Climate Change II)
Show Figures

Figure 1

Figure 1
<p>PlanetScope CNN predictions on the James H. Miller Jr. power plant at low vs. high generation on two observation dates. Separate NDT and FGD models predicted on NDT cooling tower (blue) and FGD stack (red) patches, respectively. These predictions were ingested by subsequent models to estimate generation, then CO<sub>2</sub>, for the plant. © 2023 IEEE. Reproduced with permission from [<a href="#B34-remotesensing-16-01290" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>Overview of the data and models required to estimate CO<sub>2</sub> emissions from fossil-fuel power plants using satellite imagery.</p>
Full article ">Figure 3
<p>The Ninghai power station in China as seen from Landsat 8 at 30 m spatial resolution <b>(left</b>), Sentinel-2 at 10 m spatial resolution (<b>center</b>), and PlanetScope at 3 m spatial resolution (<b>right</b>). These images only represent the visible bands (red, green, and blue).</p>
Full article ">Figure 4
<p>Overview of modeling approaches: NDT cooling towers and FGD stacks were cropped from satellite imagery (red frames above) and fed into ML models to classify the power plant’s on/off status and predict the capacity factor. Patch size is selected as detailed in <a href="#sec3dot5-remotesensing-16-01290" class="html-sec">Section 3.5</a> and may be larger than visualized here. Image from [<a href="#B33-remotesensing-16-01290" class="html-bibr">33</a>].</p>
Full article ">Figure 5
<p>A sample of model predictions for Bełchatów in Poland, Europe’s largest fossil-fuel burning power plant, for the years 2019–2022. Included are predictions from the NDT (blue line) and FGD (orange line) generation models to compare to the reported electricity generation (green line). The rolling average 30-day capacity factor never fell below 45% capacity from 2019 to 2022 and was maintained above 60% for the entirety of 2022. The NDT model performs over two times better than the FGD model for this power plant, a representative characteristic of the NDT task given its clearer signal.</p>
Full article ">Figure 6
<p>A map displaying global CO<sub>2</sub> emissions estimates produced by our ML models (equivalent to “all inference” in <a href="#remotesensing-16-01290-t002" class="html-table">Table 2</a>). Each dot represents a single power plant, with the size of the dot corresponding to the amount of estimated CO<sub>2</sub> emissions for that plant in 2022.</p>
Full article ">Figure 7
<p>A comparison for 2019–2022 of ML estimated vs. reported (<b>a</b>) capacity factor with Pearson correlation 0.80, (<b>b</b>) electricity generation (terawatt hour, TWh) with Pearson correlation 0.93, and (<b>c</b>) CO<sub>2</sub> emissions (megatonnes, Mt) with Pearson correlation 0.90 [<a href="#B50-remotesensing-16-01290" class="html-bibr">50</a>]. Each dot represents an individual plant and year matched to a reported electricity generation or emissions source from the respective region’s reporting agency. Note: reported CO<sub>2</sub> emissions were not plotted for Türkiye as no emission data are available (<a href="#secAdot5-remotesensing-16-01290" class="html-sec">Appendix A.5</a>).</p>
Full article ">Figure A1
<p>White bluff power station as shown on OpenStreetMap (<b>top</b>) and in OpenStreetMap edit mode (<b>bottom</b>). We used this aerial imagery to annotate the locations of FGD flue stacks (translucent white circle) and NDT cooling towers (red circles).</p>
Full article ">Figure A2
<p>A comparison of reported annual electricity generation (terawatt hour, TWh) and reported CO<sub>2</sub> emissions (megatonnes, Mt) of US plants (reported by CAMPD) for 2019–2022 of gas combined cycle plants with Person correlation 0.98 and coal steam turbine plants with a Pearson correlation of 0.99. Each dot represents an individual plant and year.</p>
Full article ">
22 pages, 7541 KiB  
Article
Sparse SAR Imaging Algorithm in Marine Environments Based on Memory-Augmented Deep Unfolding Network
by Yao Zhao, Chengwen Ou, He Tian, Bingo Wing-Kuen Ling, Ye Tian and Zhe Zhang
Remote Sens. 2024, 16(7), 1289; https://doi.org/10.3390/rs16071289 - 5 Apr 2024
Viewed by 1153
Abstract
Oceanic targets, including ripples, islands, vessels, and coastlines, display distinct sparse characteristics, rendering the ocean a significant arena for sparse Synthetic Aperture Radar (SAR) imaging rooted in sparse signal processing. Deep neural networks (DNNs), a current research emphasis, have, when integrated with sparse [...] Read more.
Oceanic targets, including ripples, islands, vessels, and coastlines, display distinct sparse characteristics, rendering the ocean a significant arena for sparse Synthetic Aperture Radar (SAR) imaging rooted in sparse signal processing. Deep neural networks (DNNs), a current research emphasis, have, when integrated with sparse SAR, attracted notable attention for their exceptional imaging capabilities and high computational efficiency. Yet, the efficiency of traditional unfolding techniques is impeded by their architecturally inefficient design, which curtails their information transmission capacity and consequently detracts from the quality of reconstruction. This paper unveils a novel Memory-Augmented Deep Unfolding Network (MADUN) for SAR imaging in marine environments. Our methodology harnesses the synergies between deep learning and algorithmic unfolding, enhanced with a memory component, to elevate SAR imaging’s computational precision. At the heart of our investigation is the incorporation of High-Throughput Short-Term Memory (HSM) and Cross-Stage Long-Term Memory (CLM) within the MADUN framework, ensuring robust information flow across unfolding stages and solidifying the foundation for deep, long-term informational correlations. Our experimental results demonstrate that our strategy significantly surpasses existing methods in enhancing the reconstruction of sparse marine scenes. Full article
(This article belongs to the Special Issue Radar Signal Processing and Imaging for Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Illustration of the ISTA-Net+ framework.</p>
Full article ">Figure 2
<p>Illustration of sparse SAR imaging algorithm based on MADUN.</p>
Full article ">Figure 3
<p>Illustration of the <span class="html-italic">k</span>-th stage in MADUN. “©” denotes concatenation along the channel dimension. “dCom” represents the decomposition of a complex number into its real and imaginary parts, while “Com” represents the inverse operation of “dCom”.</p>
Full article ">Figure 4
<p>Comparison of average PSNR at different stages between ISTA-Net+ and proposed method.</p>
Full article ">Figure 5
<p>Comparison of average PSNR at different epochs between ISTA-Net+ and proposed method.</p>
Full article ">Figure 6
<p>Imaging results for simulated data ships scene. From left to right are the results from the ISTA-based CS Imaging Algorithm, ISTA-Net+, and the Proposed Method. The first row displays results with subsampling at η = 80%, and the second row with subsampling at η = 50%.</p>
Full article ">Figure 7
<p>Performance curves of SAR image reconstruction in the ships scene by ISTA, ISTA-Net+, and the proposed method at different sampling ratios. (<b>a</b>) PSNR; (<b>b</b>) SSIM.</p>
Full article ">Figure 8
<p>Imaging results for simulated data dock scene. From left to right are the results from the ISTA-based CS Imaging Algorithm, ISTA-Net+, and the proposed method. The first row displays results with subsampling at η = 80%, and the second row with subsampling at η = 50%.</p>
Full article ">Figure 9
<p>Performance curves of SAR image reconstruction in the dock scene by ISTA, ISTA-Net+, and the proposed method at different sampling ratios. (<b>a</b>) PSNR; (<b>b</b>) SSIM.</p>
Full article ">Figure 10
<p>Performance comparison of phase reconstruction quality. (<b>a</b>) ISTA; (<b>b</b>) ISTA-Net+; (<b>c</b>) proposed method.</p>
Full article ">Figure 11
<p>Performance comparison of the memory enhancement mechanism. (<b>a</b>) PSNR; (<b>b</b>) SSIM; (<b>c</b>) NMSE.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) The imaging results for three different scenes based on measured data. Within the same scene, from left to right are the results yielded by the ISTA algorithm, ISTA-Net+, and the proposed method. The first row shows the results with Subsampling at η = 80%, and the second row is with subsampling at η = 50%. (<b>a</b>) Scene 1: ship scene. (<b>b</b>) Scene 2: island scene. (<b>c</b>) Scene 3: wave scene.</p>
Full article ">Figure 12 Cont.
<p>(<b>a</b>–<b>c</b>) The imaging results for three different scenes based on measured data. Within the same scene, from left to right are the results yielded by the ISTA algorithm, ISTA-Net+, and the proposed method. The first row shows the results with Subsampling at η = 80%, and the second row is with subsampling at η = 50%. (<b>a</b>) Scene 1: ship scene. (<b>b</b>) Scene 2: island scene. (<b>c</b>) Scene 3: wave scene.</p>
Full article ">
24 pages, 39812 KiB  
Article
Detection of Military Targets on Ground and Sea by UAVs with Low-Altitude Oblique Perspective
by Bohan Zeng, Shan Gao, Yuelei Xu, Zhaoxiang Zhang, Fan Li and Chenghang Wang
Remote Sens. 2024, 16(7), 1288; https://doi.org/10.3390/rs16071288 - 5 Apr 2024
Viewed by 1307
Abstract
Small-scale low-altitude unmanned aerial vehicles (UAVs) equipped with perception capability for military targets will become increasingly essential for strategic reconnaissance and stationary patrols in the future. To respond to challenges such as complex terrain and weather variations, as well as the deception and [...] Read more.
Small-scale low-altitude unmanned aerial vehicles (UAVs) equipped with perception capability for military targets will become increasingly essential for strategic reconnaissance and stationary patrols in the future. To respond to challenges such as complex terrain and weather variations, as well as the deception and camouflage of military targets, this paper proposes a hybrid detection model that combines Convolutional Neural Network (CNN) and Transformer architecture in a decoupled manner. The proposed detector consists of the C-branch and the T-branch. In the C-branch, Multi-gradient Path Network (MgpNet) is introduced, inspired by the multi-gradient flow strategy, excelling in capturing the local feature information of an image. In the T-branch, RPFormer, a Region–Pixel two-stage attention mechanism, is proposed to aggregate the global feature information of the whole image. A feature fusion strategy is proposed to merge the feature layers of the two branches, further improving the detection accuracy. Furthermore, to better simulate real UAVs’ reconnaissance environments, we construct a dataset of military targets in complex environments captured from an oblique perspective to evaluate the proposed detector. In ablation experiments, different fusion methods are validated, and the results demonstrate the effectiveness of the proposed fusion strategy. In comparative experiments, the proposed detector outperforms most advanced general detectors. Full article
Show Figures

Figure 1

Figure 1
<p>Framework of the proposed CNN–Transformer hybrid detection model.</p>
Full article ">Figure 2
<p>Region–Pixel two-stage attention mechanism of RPFormer block. The top and bottom parts are Region-stage attention and Pixel-stage attention steps, respectively. The red stars in Region-stage attention step represent the areas with the highest relevance to be retained. The arrows indicate transferring the indices of these areas to Pixel-stage attention step for sparse attention calculation.</p>
Full article ">Figure 3
<p>(<b>a</b>) The overall architecture of the RPFormer; (<b>b</b>) details of a RPFormer block.</p>
Full article ">Figure 4
<p>The efficient layer aggregation network’s (Elan) block details.</p>
Full article ">Figure 5
<p>Gradient Source of main branch in Elan block.</p>
Full article ">Figure 6
<p>(<b>a</b>) The structure of the MgpNet; (<b>b</b>) downsampling module.</p>
Full article ">Figure 7
<p>An overview structure of proposed feature fusion strategy.</p>
Full article ">Figure 8
<p>Examples of created dataset images. (<b>a</b>) Air-to-ground scenario; (<b>b</b>) air-to-sea scenario.</p>
Full article ">Figure 9
<p>Distribution details of the dataset. (<b>a</b>) Air-to-ground scenario; (<b>b</b>) air-to-sea scenario.</p>
Full article ">Figure 10
<p>The comparative inference results of different methods in the air-to-ground scenario. (<b>a</b>) Ground truth; (<b>b</b>) proposed hybrid method; (<b>c</b>) CSWin Transformer-small; (<b>d</b>) Swin Transformer-small; (<b>e</b>) PVT-small; (<b>f</b>) RestNest-50; (<b>g</b>) CspDarkNet-53; (<b>h</b>) ResNet-50; (<b>i</b>) EfficientNet-b0.</p>
Full article ">Figure 11
<p>The comparative inference results of different methods in the air-to-sea scenario. (<b>a</b>) Ground truth; (<b>b</b>) proposed hybrid method; (<b>c</b>) CSWin Transformer-small; (<b>d</b>) Swin Transformer-small; (<b>e</b>) PVT-small; (<b>f</b>) RestNest-50; (<b>g</b>) CspDarkNet-53; (<b>h</b>) ResNet-50; (<b>i</b>) EfficientNet-b0.</p>
Full article ">Figure 12
<p>The comparative feature extraction results of different methods in the air-to-ground scenario. (<b>a</b>) CspDarkNet-53 stage1; (<b>b</b>) CSWin Transformer-small stage1; (<b>c</b>) proposed hybrid method stage1; (<b>d</b>) CspDarkNet-53 stage2; (<b>e</b>) CSWin Transformer-small; (<b>f</b>) proposed hybrid method stage1; (<b>g</b>) CspDarkNet-53 stage3; (<b>h</b>) CSWin Transformer-small; (<b>i</b>) proposed hybrid method stage1.</p>
Full article ">Figure 13
<p>The comparative feature extraction results of different methods in the air-to-sea scenario. (<b>a</b>) CspDarkNet-53 stage1; (<b>b</b>) CSWin Transformer-small stage1; (<b>c</b>) proposed hybrid method stage1; (<b>d</b>) CspDarkNet-53 stage2; (<b>e</b>) CSWin Transformer-small; (<b>f</b>) proposed hybrid method stage1; (<b>g</b>) CspDarkNet-53 stage3; (<b>h</b>) CSWin Transformer-small; (<b>i</b>) proposed hybrid method stage1.</p>
Full article ">
22 pages, 15570 KiB  
Article
Time-Series Cross-Radiometric Calibration and Validation of GF-6/WFV Using Multi-Site
by Yingxian Wang, Yaokai Liu, Weiwei Zhao, Jian Zeng, Huixian Wang, Renfei Wang, Zhaopeng Xu and Qijin Han
Remote Sens. 2024, 16(7), 1287; https://doi.org/10.3390/rs16071287 - 5 Apr 2024
Cited by 1 | Viewed by 951
Abstract
The GaoFen6 (GF-6) satellite, equipped with a wide full-swath (WFV) sensor, offers high spatial resolution and extensive coverage, making it widely utilized in agricultural and forestry classification, land resource monitoring, and other fields. Accurate on-orbit radiometric calibration of GF-6/WFV is crucial for these [...] Read more.
The GaoFen6 (GF-6) satellite, equipped with a wide full-swath (WFV) sensor, offers high spatial resolution and extensive coverage, making it widely utilized in agricultural and forestry classification, land resource monitoring, and other fields. Accurate on-orbit radiometric calibration of GF-6/WFV is crucial for these quantitative applications. Currently, the absolute radiometric calibration of GF-6/WFV relies primarily on vicarious calibration conducted by the China Center for Resources Satellite Data and Application (CRESDA). However, annual vicarious calibration may not adequately capture the radiometric performance of GF-6/WFV due to performance degradation. Therefore, increasing the frequency of on-orbit radiometric calibration throughout the lifetime of GF-6/WFV is essential. This study proposes a method for conducting long-term cross-radiometric calibrations of GF-6/WFV by taking the multispectral imager (MSI) onboard the Sentinel-2 satellite as a reliable reference sensor and the sites from RadCalNet as reference ground targets. Firstly, we conducted 62 on-orbit cross-radiometric calibrations of GF-6/WFV since its launch by tracking with the Sentinel-2/MSI sensor after correcting the discrepancy spectrum and solar zenith angle. Then, validation of cross-radiometric calibration results against RadCalNet products indicated an average absolute relative error between 3.55% and 4.64%. Cross-validation with additional reference sensors, including Landsat-8/OLI and MODIS, confirmed the reliability of calibration, demonstrating relative differences from GF-6/WFV of less than 5%. Furthermore, the overall uncertainty of the cross-radiometric calibration was estimated to be from 4.08% to 4.89%. Finally, trend analysis of the time-series radiometric performance was also conducted and revealed an annual degradation rate ranging from 0.57% to 2.31%. This degradation affects surface reflectance retrieval, introducing a bias of approximately 0.0073 to 0.0084. Our findings highlight the operational effectiveness of the proposed method in achieving long-time-series on-orbit radiometric calibration and degradation monitoring of GF-6/WFV. The study also demonstrates that the radiometric performance of GF-6/WFV is relatively stable and suitable for further quantitative applications, especially for long-term monitoring applications. Full article
(This article belongs to the Special Issue Remote Sensing Satellites Calibration and Validation)
Show Figures

Figure 1

Figure 1
<p>True-color composite image of GF-6/WFV over the Gobabeb site on 4 April 2021 with the red box indicated the location of the Gobabeb site’s core region.</p>
Full article ">Figure 2
<p>True-color composite image of GF-6/WFV over the La Crau site on 24 April 2021 with the red box indicated the location of the La Crau site’s core region.</p>
Full article ">Figure 3
<p>True-color composite image of GF-6/WFV over the RVP site on 3 April 2021 with the red box indicated the location of the RVP site’s core region.</p>
Full article ">Figure 4
<p>True-color composite image of GF-6/WFV over the Baotou site on 10 September 2020 with the red box indicated the location of the Baotou Sandy site.</p>
Full article ">Figure 5
<p>Spectral response functions for matched bands of Gf-6/WFV and Sentinel-2/MSI.</p>
Full article ">Figure 6
<p>Viewing geometries of Sentinel-2/MSI and GF-6/WFV: (<b>a</b>) viewing zenith angles of Sentinel-2/MSI, (<b>b</b>) viewing zenith angles of GF-6/WFV, (<b>c</b>) relative azimuth angles of GF-6/WFV, and (<b>d</b>) relative azimuth angles of Sentinel-2/MSI.</p>
Full article ">Figure 6 Cont.
<p>Viewing geometries of Sentinel-2/MSI and GF-6/WFV: (<b>a</b>) viewing zenith angles of Sentinel-2/MSI, (<b>b</b>) viewing zenith angles of GF-6/WFV, (<b>c</b>) relative azimuth angles of GF-6/WFV, and (<b>d</b>) relative azimuth angles of Sentinel-2/MSI.</p>
Full article ">Figure 7
<p>The aerosol optical thickness at 550 nm and columnar water vapor at the time of the GF-6 satellite overpassing RadCalNet sites.</p>
Full article ">Figure 8
<p>Flowchart of the time-series cross-radiometric calibration and validation of GF-6/WFV at RadCalNet sites.</p>
Full article ">Figure 9
<p>Gains of the SBAFs between GF-6/WFV and Sentinel-2/MSI (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">Figure 9 Cont.
<p>Gains of the SBAFs between GF-6/WFV and Sentinel-2/MSI (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">Figure 10
<p>Time-series radiometric cross-calibration coefficients of GF-6/WFV (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">Figure 10 Cont.
<p>Time-series radiometric cross-calibration coefficients of GF-6/WFV (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">Figure 11
<p>Boxplot analysis of the time-series cross-radiometric calibration coefficients of GF-6/WFV.</p>
Full article ">Figure 12
<p>Validation of GF-6/WFV time-series calibration results with RadCalNet datasets.</p>
Full article ">Figure 13
<p>Cross-validation of GF-6/WFV time-series calibration results with Landsat-8/OLI datasets.</p>
Full article ">Figure 14
<p>Cross-validation of GF-6/WFV time-series calibration results with MODIS datasets.</p>
Full article ">Figure 15
<p>Trend analysis of the radiometric performance of GF-6/WFV (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">Figure 15 Cont.
<p>Trend analysis of the radiometric performance of GF-6/WFV (<b>a</b>) in the blue band, (<b>b</b>) in the green band, (<b>c</b>) in the red band, (<b>d</b>) and in the near-infrared band.</p>
Full article ">
22 pages, 15350 KiB  
Article
Spatiotemporal Analysis of Land Surface Temperature in Response to Land Use and Land Cover Changes: A Remote Sensing Approach
by Gulam Mohiuddin and Jan-Peter Mund
Remote Sens. 2024, 16(7), 1286; https://doi.org/10.3390/rs16071286 - 5 Apr 2024
Viewed by 1748
Abstract
Rapid urbanisation in the global south has often introduced substantial and rapid uncontrolled Land Use and Land Cover (LULC) changes, considerably affecting the Land Surface Temperature (LST) patterns. Understanding the relationship between LULC changes and LST is essential to mitigate such effects, considering [...] Read more.
Rapid urbanisation in the global south has often introduced substantial and rapid uncontrolled Land Use and Land Cover (LULC) changes, considerably affecting the Land Surface Temperature (LST) patterns. Understanding the relationship between LULC changes and LST is essential to mitigate such effects, considering the urban heat island (UHI). This study aims to elucidate the spatiotemporal variations and alterations of LST in urban areas compared to LULC changes. The study focused on a peripheral urban area of Phnom Penh (Cambodia) undergoing rapid urban development. Using Landsat images from 2000 to 2021, the analysis employed an exploratory time-series analysis of LST. The study revealed a noticeable variability in LST (20 to 69 °C), which was predominantly influenced by seasonal variability and LULC changes. The study also provided insights into how LST varies within different LULC at the exact spatial locations. These changes in LST did not manifest uniformly but displayed site-specific responses to LULC changes. This study accounts for changing land surfaces’ complex physical energy interaction over time. The methodology offers a replicable model for other similarly structured, rapidly urbanised regions utilising novel semi-automatic processing of LST from Landsat images, potentially inspiring future research in various urban planning and monitoring contexts. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area. (data sources: open street map, USGS for background map).</p>
Full article ">Figure 2
<p>Different Landsat satellite-wise number of images.</p>
Full article ">Figure 3
<p>Distribution of selected 425 images based on the day of the year and Landsat satellite type.</p>
Full article ">Figure 4
<p>Graphical overview of the methodical overview.</p>
Full article ">Figure 5
<p>Boxplot of minimum, mean and maximum LST (number of observation (N) = 425).</p>
Full article ">Figure 6
<p>Calendar heatmap of mean LST (white boxes inside the figure represent no data).</p>
Full article ">Figure 7
<p>Boxplot of yearly minimum, mean and maximum LST.</p>
Full article ">Figure 8
<p>Total time series of minimum, mean and maximum LST (2000–2021).</p>
Full article ">Figure 9
<p>LST (in °C) in different months from 2015 (images in March, July, August and November were heavily contaminated from cloud).</p>
Full article ">Figure 10
<p>Monthly minimum, mean and maximum LST in 2015.</p>
Full article ">Figure 11
<p>Example of how visual interpretation provides additional input.</p>
Full article ">Figure 12
<p>LST (in °C) in 2000–2021.</p>
Full article ">Figure 13
<p>Visible connection between LST (in °C) and LULC changes elaborating spatial changes: (<b>a</b>) LST in March 2003; (<b>b</b>) Google Earth RGB image in February 2003; (<b>c</b>) LST in March 2019; (<b>d</b>) Google Earth RGB image in February 2019.</p>
Full article ">Figure 14
<p>LST and LULC changes in point 4 (The blue line is the line graph of LST data in the time series, and the green line is a statistical trendline showing the upward trend of LST at this specific point).</p>
Full article ">Figure 15
<p>Correlation matrix between LST, IBI, MNDWI and NDVI in 400 random points.</p>
Full article ">Figure 16
<p>Linear relationship between built-up area and LST (in °C).</p>
Full article ">Figure 17
<p>Relationship between LST and NDVI and MNDWI: (<b>a</b>) LST vs. NDVI with all 400 samples; (<b>b</b>) LST vs. NDVI with only points that have ≥ 0.2 NDVI values (only vegetation samples); (<b>c</b>) LST vs. MNDWI with all 400 samples; (<b>d</b>) LST vs. MNDWI with only points that has ≤0.2 NDVI (without vegetation samples).</p>
Full article ">Figure A1
<p>Location 400 random points considered for correlation test between LST and LULC.</p>
Full article ">Figure A2
<p>Location of identified five points to examine LST and LULC interactions (background map is LST (in °C) from April 2015).</p>
Full article ">
25 pages, 7926 KiB  
Article
Generative Adversarial Network and Mutual-Point Learning Algorithm for Few-Shot Open-Set Classification of Hyperspectral Images
by Tuo Xu, Ying Wang, Jie Li and Yuefan Du
Remote Sens. 2024, 16(7), 1285; https://doi.org/10.3390/rs16071285 - 5 Apr 2024
Viewed by 923
Abstract
Existing approaches addressing the few-shot open-set recognition (FSOSR) challenge in hyperspectral images (HSIs) often encounter limitations stemming from sparse labels, restricted category numbers, and low openness. These limitations compromise stability and adaptability. In response, an open-set HSI classification algorithm based on data wandering [...] Read more.
Existing approaches addressing the few-shot open-set recognition (FSOSR) challenge in hyperspectral images (HSIs) often encounter limitations stemming from sparse labels, restricted category numbers, and low openness. These limitations compromise stability and adaptability. In response, an open-set HSI classification algorithm based on data wandering (DW) is introduced in this research. Firstly, a K-class classifier suitable for a closed set is trained, and its internal encoder is leveraged to extract features and estimate the distribution of known categories. Subsequently, the classifier is fine-tuned based on feature distribution. To address the scarcity of samples, a sample density constraint based on the generative adversarial network (GAN) is employed to generate synthetic samples near the decision boundary. Simultaneously, a mutual-point learning method is incorporated to widen the class distance between known and unknown categories. In addition, a dynamic threshold method based on DW is devised to enhance the open-set performance. By categorizing drifting synthetic samples into known and unknown classes and retraining them together with the known samples, the closed-set classifier is optimized, and a (K + 1)-class open-set classifier is trained. The experimental results in this research demonstrate the superior FSOSR performance of the proposed method across three benchmark HSI datasets. Full article
Show Figures

Figure 1

Figure 1
<p>Generating samples near the decision boundary.</p>
Full article ">Figure 2
<p>Overall network architecture.</p>
Full article ">Figure 3
<p>Pseudo-sample center.</p>
Full article ">Figure 4
<p>Graphical representation of RPL. (<b>a</b>) Reciprocal points of a known category. (<b>b</b>) Mutual exclusion between all known categories and reciprocal points. Where ★ signifies the open-set space of unknown samples, ▲ symbolizes the reciprocal points, and ● marks the known samples.</p>
Full article ">Figure 5
<p>A learnable dynamic thresholding strategy for open space, (<b>a</b>) schematic of RPL, (<b>b</b>) reinforced boundaries via GAN, and (<b>c</b>) high-dimensional feature space zoomed in via DW, where the blue region marks the constructed separation space.</p>
Full article ">Figure 6
<p>Analysis combined with data walk sample distribution.</p>
Full article ">Figure 7
<p>Sample distribution and its information energy field before data walking.</p>
Full article ">Figure 8
<p>Sample distribution and its information energy field after six iterations.</p>
Full article ">Figure 9
<p>PU dataset. (<b>a</b>) False-color image. (<b>b</b>) Closed-set ground truth.</p>
Full article ">Figure 10
<p>SA dataset. (<b>a</b>) False-color image. (<b>b</b>) Closed-set ground truth.</p>
Full article ">Figure 11
<p>IP dataset. (<b>a</b>) False-color image. (<b>b</b>) Closed-set ground truth.</p>
Full article ">Figure 12
<p>Thematic classification maps of different methods. (<b>a</b>) Ground truth map. (<b>b</b>) SSTN. (<b>c</b>) SnaTCHer. (<b>d</b>) PEELER. (<b>e</b>) RDOSR. (<b>f</b>) OpenMax. (<b>g</b>) MDL4OW. (<b>h</b>) SSLR. (<b>i</b>) RPLDW.</p>
Full article ">Figure 13
<p>T-SNE visualization results of feature space on SA data. (<b>a</b>) Original feature space. (<b>b</b>) Feature space constructed by RPLDW (the 16th category is unknown). (<b>c</b>) Feature space constructed by RPLDW (the 11th, 12th, 13th, and 14th categories are defined as the unknown ones).</p>
Full article ">
17 pages, 6317 KiB  
Article
Spectral Reconstruction from Thermal Infrared Multispectral Image Using Convolutional Neural Network and Transformer Joint Network
by Enyu Zhao, Nianxin Qu, Yulei Wang and Caixia Gao
Remote Sens. 2024, 16(7), 1284; https://doi.org/10.3390/rs16071284 - 5 Apr 2024
Cited by 1 | Viewed by 1084
Abstract
Thermal infrared remotely sensed data, by capturing the thermal radiation characteristics emitted by the Earth’s surface, plays a pivotal role in various domains, such as environmental monitoring, resource exploration, agricultural assessment, and disaster early warning. However, the acquisition of thermal infrared hyperspectral remotely [...] Read more.
Thermal infrared remotely sensed data, by capturing the thermal radiation characteristics emitted by the Earth’s surface, plays a pivotal role in various domains, such as environmental monitoring, resource exploration, agricultural assessment, and disaster early warning. However, the acquisition of thermal infrared hyperspectral remotely sensed imagery necessitates more complex and higher-precision sensors, which in turn leads to higher research and operational costs. In this study, a novel Convolutional Neural Network (CNN)–Transformer combined block, termed CTBNet, is proposed to address the challenge of thermal infrared multispectral image spectral reconstruction. Specifically, the CTBNet comprises blocks that integrate CNN and Transformer technologies (CTB). Within these CTBs, an improved self-attention mechanism is introduced, which not only considers features across spatial and spectral dimensions concurrently, but also explicitly extracts incremental features from each channel. Compared to other algorithms, the proposed method more closely aligns with the true spectral curves in the reconstruction of hyperspectral images across the spectral dimension. Through a series of experiments, this approach has been proven to ensure robustness and generalizability, outperforming some state-of-the-art algorithms across various metrics. Full article
Show Figures

Figure 1

Figure 1
<p>CTBNet structure.</p>
Full article ">Figure 2
<p>CTB structure. (<b>a</b>) overall structure of CTB; (<b>b</b>) composition of FNN; (<b>c</b>) process of K Map; (<b>d</b>) process of V Map.</p>
Full article ">Figure 3
<p>A portion of the hyperspectral data: (<b>a</b>) including various man-made structures; (<b>b</b>) encompassing rivers, vegetation, highways, and other elements.</p>
Full article ">Figure 4
<p>Simulated spectral response functions.</p>
Full article ">Figure 5
<p>Comparison of results for land region. (<b>a</b>) Error map for land region. (<b>b</b>) Truth map for the land region.</p>
Full article ">Figure 5 Cont.
<p>Comparison of results for land region. (<b>a</b>) Error map for land region. (<b>b</b>) Truth map for the land region.</p>
Full article ">Figure 6
<p>Comparison of results for water region. (<b>a</b>) Error map for the water region (<b>b</b>) True value for the water region.</p>
Full article ">Figure 6 Cont.
<p>Comparison of results for water region. (<b>a</b>) Error map for the water region (<b>b</b>) True value for the water region.</p>
Full article ">Figure 7
<p>Radiance spectral curve. (<b>a</b>) for the land region and (<b>b</b>) for the water region.</p>
Full article ">Figure 8
<p>Spectral response function of simulated MODIS.</p>
Full article ">
23 pages, 7359 KiB  
Article
Spatiotemporal Variation and Causes of Typical Extreme Precipitation Events in Shandong Province over the Last 50 Years
by Jie Liu, Baofu Li and Mengqiu Ma
Remote Sens. 2024, 16(7), 1283; https://doi.org/10.3390/rs16071283 - 5 Apr 2024
Cited by 1 | Viewed by 809
Abstract
In this study, based on hourly ERA5 reanalysis data from July to September, from 1971 to 2020, for Shandong Province, we used mathematical statistical analysis, the Mann–Kendall nonparametric statistical test, cluster analysis, and other methods to extract and analyze the spatiotemporal variation characteristics [...] Read more.
In this study, based on hourly ERA5 reanalysis data from July to September, from 1971 to 2020, for Shandong Province, we used mathematical statistical analysis, the Mann–Kendall nonparametric statistical test, cluster analysis, and other methods to extract and analyze the spatiotemporal variation characteristics and causes of typical extreme precipitation events. The results indicated the following: (1) The total number and duration of precipitation events show a nonsignificant upward trend, while the average and extreme rainfall intensities show a nonsignificant downward trend. (2) Extreme precipitation events are primarily concentrated in Qingdao, Jinan, Heze, and Binzhou, with fewer events occurring in central Shandong Province. (3) Extreme precipitation events are classified into four types (namely, patterns I, II, III, and IV). Pattern I exhibits two rain peaks, with the primary rain peak occurring after the secondary rain peak. Similarly, pattern II also displays two rain peaks, with equivalent rainfall amounts for both peaks. In contrast, pattern III has multiple, evenly distributed rain peaks. Finally, pattern IV shows a rain peak during the first half of the precipitation event. Pattern I has the highest occurrence probability (46%), while pattern IV has the lowest (7%). (4) The spatial distributions of the different rain patterns are similar, with most being found in the eastern coastal and western regions. (5) Extreme precipitation events result from interactions between large-scale circulation configurations and mesoscale convective systems. The strong blocking situation and significant circulation transport at middle and low latitudes in East Asia, along with strong convergent uplift, abnormally high specific humidity, and high-water-vapor convergence centers, play crucial roles in supporting large-scale circulation systems and triggering mesoscale convective systems. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location map of Shandong Province.</p>
Full article ">Figure 2
<p>Typical extreme precipitation total amount trend map (<b>a</b>) and spatial distribution map (<b>b</b>) of Shandong Province (1971–2020).</p>
Full article ">Figure 3
<p>The typical extreme precipitation duration trend (<b>a</b>) and spatial distribution (<b>b</b>) in Shandong Province (1971–2020).</p>
Full article ">Figure 4
<p>Typical extreme rainfall intensity trend map (<b>a</b>), average rainfall intensity spatial distribution map (<b>b</b>), and extreme rainfall intensity spatial distribution map (<b>c</b>) of Shandong Province (1971–2020).</p>
Full article ">Figure 5
<p>Statistical chart of the typical extreme precipitation start time (<b>a</b>), statistical chart of the rainstorm start time (<b>b</b>), spatial distribution map of the precipitation start time (<b>c</b>), and spatial distribution map of the rainstorm start time (<b>d</b>) in Shandong Province (1971–2020).</p>
Full article ">Figure 6
<p>Typical extreme precipitation frequency distribution in Shandong Province (1971–2020).</p>
Full article ">Figure 7
<p>Classification of the different rain patterns of typical extreme precipitation events in Shandong Province from 1971 to 2020 ((<b>a</b>) rain pattern I, (<b>b</b>) rain pattern II, (<b>c</b>) rain pattern III, and (<b>d</b>) rain pattern IV; color line: precipitation process of extreme precipitation events).</p>
Full article ">Figure 8
<p>Typical extreme precipitation frequency distributions of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) rain pattern I, (<b>b</b>) rain pattern II, (<b>c</b>) rain pattern III, and (<b>d</b>) rain pattern IV).</p>
Full article ">Figure 8 Cont.
<p>Typical extreme precipitation frequency distributions of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) rain pattern I, (<b>b</b>) rain pattern II, (<b>c</b>) rain pattern III, and (<b>d</b>) rain pattern IV).</p>
Full article ">Figure 9
<p>Statistical chart of the typical extreme precipitation in the year of occurrence (<b>a</b>) and monthly time of occurrence (<b>b</b>) for the different rain patterns in Shandong Province from 1971 to 2020.</p>
Full article ">Figure 10
<p>Extreme precipitation average in the field for the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) rain pattern I, (<b>b</b>) rain pattern II, (<b>c</b>) rain pattern III, and (<b>d</b>) rain pattern IV; shadow: 500 hPa height field, unit: gpm; arrow: 850 hPa vector wind field, unit: m/s; A: anticyclone, and C: cyclone).</p>
Full article ">Figure 11
<p>Typical extreme precipitation average dynamic fields of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) 200 hPa, (<b>b</b>) 700 hPa, and (<b>c</b>) 925 hPa; 1. rain pattern I, 2. rain pattern II, 3. rain pattern III, and 4. rain pattern IV; contour line: divergence, unit: s<sup>−1</sup>; shadow: vertical velocity, unit: Pa·s<sup>−1</sup>).</p>
Full article ">Figure 11 Cont.
<p>Typical extreme precipitation average dynamic fields of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) 200 hPa, (<b>b</b>) 700 hPa, and (<b>c</b>) 925 hPa; 1. rain pattern I, 2. rain pattern II, 3. rain pattern III, and 4. rain pattern IV; contour line: divergence, unit: s<sup>−1</sup>; shadow: vertical velocity, unit: Pa·s<sup>−1</sup>).</p>
Full article ">Figure 12
<p>Typical extreme precipitation average water vapor fields of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) 700 hPa, (<b>b</b>) 850 hPa, and (<b>c</b>) 925 hPa; 1. rain pattern I, 2. rain pattern II, 3. rain pattern III, and 4. rain pattern IV; contour line: specific humidity, unit: g·kg<sup>−1</sup>; shadow: water vapor flux divergence, unit: g·cm<sup>−2</sup>·hpa<sup>−1</sup>·s<sup>−1</sup>).</p>
Full article ">Figure 12 Cont.
<p>Typical extreme precipitation average water vapor fields of the different rain patterns in Shandong Province from 1971 to 2020 ((<b>a</b>) 700 hPa, (<b>b</b>) 850 hPa, and (<b>c</b>) 925 hPa; 1. rain pattern I, 2. rain pattern II, 3. rain pattern III, and 4. rain pattern IV; contour line: specific humidity, unit: g·kg<sup>−1</sup>; shadow: water vapor flux divergence, unit: g·cm<sup>−2</sup>·hpa<sup>−1</sup>·s<sup>−1</sup>).</p>
Full article ">
19 pages, 10533 KiB  
Article
Interannual Variations in the Summer Coastal Upwelling in the Northeastern South China Sea
by Wuyang Chen, Yifeng Tong, Wei Li, Yang Ding, Junmin Li, Wenhua Wang and Ping Shi
Remote Sens. 2024, 16(7), 1282; https://doi.org/10.3390/rs16071282 - 5 Apr 2024
Cited by 1 | Viewed by 783
Abstract
This study scrutinizes interannual (2003–2023) variations in coastal upwelling along the Guangdong Province during summers (June–August) in the northeastern South China Sea (NESCS) by comprehensively applying the moderate-resolution imaging spectroradiometer (MODIS) remote sensing sea surface temperature (SST) and chlorophyll concentration (CHL) data and [...] Read more.
This study scrutinizes interannual (2003–2023) variations in coastal upwelling along the Guangdong Province during summers (June–August) in the northeastern South China Sea (NESCS) by comprehensively applying the moderate-resolution imaging spectroradiometer (MODIS) remote sensing sea surface temperature (SST) and chlorophyll concentration (CHL) data and the model reanalysis product. The results show that SST and upwelling intensity in the sea area have significant (p < 0.05) rising trends in the last 21 years. The CHL shows an upward but insignificant trend, which is affected simultaneously by the rise in SST and the enhancement of upwelling. Further analysis reveals that the interannual variations in upwelling are robustly related to the wind fields’ variations in the coastal region. A clockwise/counter-clockwise anomaly in the wind field centered on the NESCS facilitates alongshore/onshore winds near the Guangdong coast, which can strengthen/weaken coastal upwelling. Based on the correlation between wind field variations and large-scale climate factors, long-term variations in the upwelling intensity can be primarily predicted by the Oceanic Niño Index. Full article
Show Figures

Figure 1

Figure 1
<p>Climatological (2003–2023) (<b>a</b>) sea surface temperature (SST), (<b>b</b>) topographic position index (TPI), (<b>c</b>) probability of TPI &lt; −0.5, and (<b>d</b>) chlorophyll concentration (CHL) in the northeastern South China Sea (NESCS) during summer (June, July, and August); the boxes in (<b>b</b>,<b>c</b>) represent the regions of the coastal upwelling and its three cores focused upon in this study.</p>
Full article ">Figure 2
<p>Scatter comparisons of summer mean SST/TPI/CHL based on monthly MODIS data versus minimum SST/minimum TPI/maximum CHL based on 8-day MODIS data for the upwelling regions in the NESCS, as marked in <a href="#remotesensing-16-01282-f001" class="html-fig">Figure 1</a>, during summers from 2003 to 2023.</p>
Full article ">Figure 3
<p>Annual variation in summer mean SST (°C), TPI (°C), and CHL (mg/m<sup>3</sup>) in the upwelling regions of the NESCS, as marked in <a href="#remotesensing-16-01282-f001" class="html-fig">Figure 1</a>. The dots with lines represent the SST/TPI/CHL, the solid lines represent the trend by Theil–Sen estimation, and the dashed lines represent the confidence intervals.</p>
Full article ">Figure 4
<p>Spatial distribution of the Sen slopes of the summer seasonal mean (<b>a</b>) SST (°C/year), (<b>b</b>) TPI (°C/year), and (<b>c</b>) CHL (mg/m<sup>3</sup>/year) in the NESCS during 2003–2023. Trends values that are statistically significant (<span class="html-italic">p</span> &lt; 0.05) according to the Mann–Kendall test are marked as black crosses.</p>
Full article ">Figure 5
<p>Climatological (2003–2023) (<b>a</b>) sea surface wind field and its (<b>b</b>) alongshore and (<b>c</b>) onshore components in the NESCS during summer. The box denotes the statistical region of <a href="#remotesensing-16-01282-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 6
<p>Multiyear series of the average wind vector and the total, alongshore, and onshore wind speeds in the NESCS during summers. The statistical region is the box in <a href="#remotesensing-16-01282-f005" class="html-fig">Figure 5</a>. The dots with lines represent the total/alongshore/onshore wind speeds, the solid lines represent the trend by Theil–Sen estimation, and the dashed lines represent the confidence intervals.</p>
Full article ">Figure 7
<p>Spatial distributions of the summer seasonal mean topographic position index (TPI) and wind field in the NESCS from 2002 to 2023.</p>
Full article ">Figure 8
<p>The summer-seasonal composite wind fields and TPI values in the NESCS during (<b>a</b>) weak and (<b>b</b>) strong upwelling years, as well as (<b>c</b>,<b>d</b>) their anomalies relative to the climatological means, and (<b>e</b>,<b>f</b>) the alongshore components of the wind anomalies. The weak (strong) upwelling years are when the summer TPI values in the UPW region are above (below) the trend line in <a href="#remotesensing-16-01282-f003" class="html-fig">Figure 3</a>.</p>
Full article ">Figure 9
<p>The series of the Oceanic Niño Index (ONI) and the summer TPI in the UPW region after deducting its long-term trend. The blue, red, and dashed lines indicate that the ONI value equals to 0.5, –0.5, and 0, respectively.</p>
Full article ">Figure 10
<p>Composite wind field anomalies in the South China Sea and Western Pacific during the summer following the (<b>left panel</b>) El Niño and (<b>right panel</b>) La Niña years.</p>
Full article ">Figure 11
<p>The correlation coefficient <span class="html-italic">r</span> between the TPI and the ONI change rate (i.e., ONI<span class="html-italic"><sub>i</sub></span>–ONI<span class="html-italic"><sub>j</sub></span>). The corresponding <span class="html-italic">p</span>-value is displayed after transposition on the top-left corner of each panel for the upwelling regions (i.e., UPW, UPW-A, UPW-B, and UPW-C marked in <a href="#remotesensing-16-01282-f001" class="html-fig">Figure 1</a>). The −11th to −6th months represent July to December of the previous years, and the −5th to −1st months represent January to May before the given summer. The stars denote the most significant correlations with the lowest <span class="html-italic">p</span>-value and its corresponding <span class="html-italic">r</span>-value for the upwelling regions.</p>
Full article ">Figure 12
<p>Comparison of the TPI series between the empirical formula (Equation (7)) and remote sensing observations for the upwelling regions (i.e., UPW, UPW-A, UPW-B, and UPW-C marked in <a href="#remotesensing-16-01282-f001" class="html-fig">Figure 1</a>).</p>
Full article ">
22 pages, 7233 KiB  
Article
High-Resolution Canopy Height Mapping: Integrating NASA’s Global Ecosystem Dynamics Investigation (GEDI) with Multi-Source Remote Sensing Data
by Cesar Alvites, Hannah O’Sullivan, Saverio Francini, Marco Marchetti, Giovanni Santopuoli, Gherardo Chirici, Bruno Lasserre, Michela Marignani and Erika Bazzato
Remote Sens. 2024, 16(7), 1281; https://doi.org/10.3390/rs16071281 - 5 Apr 2024
Cited by 2 | Viewed by 2552
Abstract
Accurate structural information about forests, including canopy heights and diameters, is crucial for quantifying tree volume, biomass, and carbon stocks, enabling effective forest ecosystem management, particularly in response to changing environmental conditions. Since late 2018, NASA’s Global Ecosystem Dynamics Investigation (GEDI) mission has [...] Read more.
Accurate structural information about forests, including canopy heights and diameters, is crucial for quantifying tree volume, biomass, and carbon stocks, enabling effective forest ecosystem management, particularly in response to changing environmental conditions. Since late 2018, NASA’s Global Ecosystem Dynamics Investigation (GEDI) mission has monitored global canopy structure using a satellite Light Detection and Ranging (LiDAR) instrument. While GEDI has collected billions of LiDAR shots across a near-global range (between 51.6°N and >51.6°S), their spatial distribution remains dispersed, posing challenges for achieving complete forest coverage. This study proposes and evaluates an approach that generates high-resolution canopy height maps by integrating GEDI data with Sentinel-1, Sentinel-2, and topographical ancillary data through three machine learning (ML) algorithms: random forests (RF), gradient tree boost (GB), and classification and regression trees (CART). To achieve this, the secondary aims included the following: (1) to assess the performance of three ML algorithms, RF, GB, and CART, in predicting canopy heights, (2) to evaluate the performance of our canopy height maps using reference canopy height from canopy height models (CHMs), and (3) to compare our canopy height maps with other two existing canopy height maps. RF and GB were the top-performing algorithms, achieving the best 13.32% and 16% root mean squared error for broadleaf and coniferous forests, respectively. Validation of the proposed approach revealed that the 100th and 98th percentile, followed by the average of the 75th, 90th, 95th, and 100th percentiles (AVG), were the most accurate GEDI metrics for predicting real canopy heights. Comparisons between predicted and reference CHMs demonstrated accurate predictions for coniferous stands (R-squared = 0.45, RMSE = 29.16%). Full article
(This article belongs to the Special Issue Vegetation Structure Monitoring with Multi-Source Remote Sensing Data)
Show Figures

Figure 1

Figure 1
<p>Study site locations and diametric class distribution of trees. (<b>a</b>) Locations visualised using a single RBG Sentinel − 2 image. (<b>b</b>) The diametric class distribution of trees in Pennataro (broadleaf stand). (<b>c</b>) The diametric class distribution of trees in Lago di Occhito (coniferous stand). The diametric class distribution of trees in Pennataro is positively skewed, indicating higher structural heterogeneity, compared to Lago di Occhito, which follows a normal distribution, indicating structural homogeneity.</p>
Full article ">Figure 2
<p>Canopy height map workflow. Random forests, gradient tree boost, and classification and regression trees were used to generate canopy height maps. The resulting canopy height map was validated using canopy height models from airborne laser scanning (ALS-based CHM) data through root mean square error (RMSE) and coefficient determination (R-squared).</p>
Full article ">Figure 3
<p>Map accuracy of downscaled canopy height maps. Five GEDI (Global Ecosystem Dynamics Investigation) metrics (Rh; Rh90, Rh95, Rh98, Rh100, and AVG of Rh75, Rh90, Rh95, Rh100) were predicted using three machine learning (ML) algorithms, namely RF, GB, and CART.</p>
Full article ">Figure 4
<p>Predicted canopy height maps for each GEDI Rh metric and ML algorithm for the coniferous forest site. A subsection has been enlarged for interpretability.</p>
Full article ">Figure 5
<p>The predicted canopy height maps for all GEDI relative height metrics and ML algorithms for the broadleaf forest site.</p>
Full article ">Figure 6
<p>Scatter plots comparing predicted and reference canopy heights in the coniferous site. Five GEDI metrics (Rh90, Rh95, Rh98, Rh100, and the average of Rh75, Rh90, Rh95, and Rh100—AVG) were processed through RF (Random Forests), GB (Gradient Tree Boost), and CART (Classification and Regression Trees) algorithms. Root mean square error (RMSE in meter and percentage) and R-squared values are presented.</p>
Full article ">Figure 7
<p>Scatter plots comparing predicted and reference canopy heights in the broadleaf site. Five GEDI metrics (Rh90, Rh95, Rh98, Rh100, and the average of Rh75, Rh90, Rh95, and Rh100—AVG) were processed through RF (Random Forests), GB (Gradient Tree Boost), and CART (Classification and Regression Trees) algorithms. Root mean square error (RMSE in meter and percentage) and R-squared values are presented.</p>
Full article ">Figure 8
<p>Scatter plot and horizontal pixel frequency distribution comparing predicted canopy heights with reference data in a coniferous forest site. Canopy heights from ALS, RF_Rh90, CH-Potapov2019, and CH-Lang2020 were utilized. Root mean square error (RMSE in m and %) and R-squared values are presented.</p>
Full article ">Figure 9
<p>Scatter plot and horizontal pixel frequency distribution comparing predicted canopy heights with reference data in a broadleaf forest site. Canopy heights from ALS, RF_Rh90, CH-Potapov2019, and CH-Lang2020 were utilized. Root mean square error (RMSE in m and %) and R-squared values are presented.</p>
Full article ">
20 pages, 15304 KiB  
Article
Detection and Attribution of Vegetation Dynamics in the Yellow River Basin Based on Long-Term Kernel NDVI Data
by Haiying Yu, Qianhua Yang, Shouzheng Jiang, Bao Zhan and Cun Zhan
Remote Sens. 2024, 16(7), 1280; https://doi.org/10.3390/rs16071280 - 5 Apr 2024
Cited by 2 | Viewed by 1384
Abstract
Detecting and attributing vegetation variations in the Yellow River Basin (YRB) is vital for adjusting ecological restoration strategies to address the possible threats posed by changing environments. On the basis of the kernel normalized difference vegetation index (kNDVI) and key climate [...] Read more.
Detecting and attributing vegetation variations in the Yellow River Basin (YRB) is vital for adjusting ecological restoration strategies to address the possible threats posed by changing environments. On the basis of the kernel normalized difference vegetation index (kNDVI) and key climate drivers (precipitation (PRE), temperature (TEM), solar radiation (SR), and potential evapotranspiration (PET)) in the basin during the period from 1982 to 2022, we utilized the multivariate statistical approach to analyze the spatiotemporal patterns of vegetation dynamics, identified the key climate variables, and discerned the respective impacts of climate change (CC) and human activities (HA) on these variations. Our analysis revealed a widespread greening trend across 93.1% of the YRB, with 83.2% exhibiting significant increases in kNDVI (p < 0.05). Conversely, 6.9% of vegetated areas displayed a browning trend, particularly concentrated in the alpine and urban areas. With the Hurst index of kNDVI exceeding 0.5 in 97.5% of vegetated areas, the YRB tends to be extensively greened in the future. Climate variability emerges as a pivotal determinant shaping diverse spatial and temporal vegetation patterns, with PRE exerting dominance in 41.9% of vegetated areas, followed by TEM (35.4%), SR (13%), and PET (9.7%). Spatially, increased PRE significantly enhanced vegetation growth in arid zones, while TEM and SR controlled vegetation variations in alpine areas and non-water-limited areas such as irrigation zones. Vegetation dynamics in the YRB were driven by a combination of CC and HA, with relative contributions of 55.8% and 44.2%, respectively, suggesting that long-term CC is the dominant force. Specifically, climate change contributed to the vegetation greening seen in the alpine region and southeastern part of the basin, and human-induced factors benefited vegetation growth on the Loess Plateau (LP) while inhibiting growth in urban and alpine pastoral areas. These findings provide critical insights that inform the formulation and adaptation of ecological conservation strategies in the basin, thereby enhancing resilience to changing environmental conditions. Full article
(This article belongs to the Special Issue Remote Sensing of Arid/Semiarid Lands II)
Show Figures

Figure 1

Figure 1
<p>Spatial patterns of the multiyear average of climate variables and vegetation types in the Yellow River Basin. (<b>a</b>) Geographic location and elevation; (<b>b</b>) precipitation; (<b>c</b>) temperature; (<b>d</b>) solar radiation; (<b>e</b>) potential evapotranspiration; (<b>f</b>) land-use types in 2000.</p>
Full article ">Figure 2
<p>Spatial patterns of (<b>a</b>) multi-year average values and (<b>b</b>) the variation coefficient of growing season <span class="html-italic">kNDVI</span> in the Yellow River Basin from 1982 to 2022.</p>
Full article ">Figure 3
<p>(<b>a</b>) Interannual slopes of the kernel normalized difference vegetation index (<span class="html-italic">kNDVI</span>); spatial distribution of (<b>b</b>) <span class="html-italic">kNDVI</span> trend, (<b>c</b>) Hurst index for <span class="html-italic">kNDVI</span>, and (<b>d</b>) future <span class="html-italic">kNDVI</span> trends in the Yellow River Basin from 1982 to 2022.</p>
Full article ">Figure 4
<p>Spatial distribution of the (<b>a</b>) determination coefficient and (<b>b</b>) areas corresponding to significant relationships (<span class="html-italic">p</span> &lt; 0.05) between the kernel normalized difference vegetation index (<span class="html-italic">kNDVI</span>) and the four main climate factors identified by the multiple regression model.</p>
Full article ">Figure 5
<p>Spatial patterns of partial correlation coefficients (PCC) between the kernel normalized difference vegetation index (<span class="html-italic">kNDVI</span>) and (<b>a</b>) precipitation, (<b>b</b>) temperature, (<b>c</b>) solar radiation, and (<b>d</b>) potential evapotranspiration. (<b>e</b>) Spatial patterns of the climate drivers dominating vegetation dynamics; (<b>f</b>) proportion of the controlled area of dominant climate drivers. Black dots represent significant correlations (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Spatial patterns of the impacts of (<b>a</b>) climate change and (<b>b</b>) human activities on vegetation dynamics in the Yellow River Basin from 1982 to 2022.</p>
Full article ">Figure 7
<p>Spatial distribution of the driver classifications of vegetation dynamics in the Yellow River Basin from 1982 to 2022. CC and HA denote climate change and human activities, respectively.</p>
Full article ">Figure 8
<p>Spatial distribution of the relative contributions of climatic change (<b>a</b>) and human activities (<b>b</b>) to vegetation dynamics in the Yellow River Basin from 1982 to 2022.</p>
Full article ">Figure 9
<p>Spatial distribution of land-use types in (<b>a</b>) 1980, (<b>b</b>) 2000, and (<b>c</b>) 2020, and land-use conversion in (<b>d</b>) 1980–2000, (<b>e</b>) 2000–2020, and (<b>f</b>) 1980–2020 in the Yellow River Basin (unit: km<sup>2</sup>). The colors in subfigures (<b>d</b>–<b>f</b>) are consistent with subfigures (<b>a</b>–<b>c</b>), representing different land use types.</p>
Full article ">Figure 10
<p>Satellite images of major cities in the Yellow River Basin.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop