[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (200)

Search Parameters:
Keywords = spaceborne lidar

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 3939 KiB  
Article
Combining LiDAR, SAR, and DEM Data for Estimating Understory Terrain Using Machine Learning-Based Methods
by Jiapeng Huang, Yue Zhang and Jianhuang Ding
Forests 2024, 15(11), 1992; https://doi.org/10.3390/f15111992 - 11 Nov 2024
Viewed by 729
Abstract
Currently, precise estimation of understory terrain faces numerous technical obstacles and challenges that are difficult to overcome. To address this problem, this paper combines LiDAR, SAR, and DEM data to estimate understory terrain. The high multivariable-precision spaceborne LiDAR ICESat-2 data, validated by the [...] Read more.
Currently, precise estimation of understory terrain faces numerous technical obstacles and challenges that are difficult to overcome. To address this problem, this paper combines LiDAR, SAR, and DEM data to estimate understory terrain. The high multivariable-precision spaceborne LiDAR ICESat-2 data, validated by the NEON, are divided into training and validation sets. The training dataset is used as a dependent variable, the SRTM DEM and Sentinel-1 SAR data are regarded as independent variables, a total of 13 feature parameters with high contributions are extracted to construct a Multiple Linear Regression model (MLR), BAGGING model, Random Forest model (RF), and Long Short-Term Memory model (LSTM). The results indicate that the RF model exhibits the highest accuracy among the four models, with R2 = 0.999, RMSE = 0.701 m, and MAE = 0.249 m. Then, based on the RF model, the understory terrain at the regional scale is generated, and an accuracy assessment is performed using the validation dataset, yielding R2 = 0.999, RMSE = 0.847 m, and MAE = 0.517 m. Furthermore, this paper quantitatively analyzes the effects of slope, vegetation coverage, and canopy height on the estimation accuracy of understory terrain. The results show that as slope, and canopy height increase, the estimation accuracy of the RF model for understory terrain gradually decreases. The accuracy of the understory terrain estimated by the RF model is relatively stable and not easily affected by slope, vegetation coverage, and canopy height. The research on the estimation of understory terrain holds significant practical implications for forest resource management, ecological conservation, and biodiversity protection, as well as natural disaster prevention. Full article
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the SRTM DEM data for the study area.</p>
Full article ">Figure 2
<p>The flowchart of the method for estimating understory terrain using the LiDAR and SAR data.</p>
Full article ">Figure 3
<p>The schematic diagram of the RF model.</p>
Full article ">Figure 4
<p>The scatter plot of the correlation between the NEON DTM data and ICESat-2 understory terrain data.</p>
Full article ">Figure 5
<p>The schematic diagram of the regional-scale estimation using the RF Model.</p>
Full article ">Figure 6
<p>The schematic diagram of the SHAP values for characteristic variables.</p>
Full article ">Figure 7
<p>The schematic diagram of the evaluation indicators of the accuracy of the regional-scale data obtained by the SRTM and RF models under different slope conditions: (<b>a</b>) R<sup>2</sup>; (<b>b</b>) RMSE.</p>
Full article ">Figure 8
<p>The schematic diagram of the evaluation indicators of the accuracy of the regional-scale data obtained by the SRTM and RF models under different vegetation coverage conditions: (<b>a</b>) R<sup>2</sup>; (<b>b</b>) RMSE.</p>
Full article ">Figure 9
<p>The schematic diagram of the evaluation indicators of the accuracy of the regional-scale data obtained by the SRTM and RF models under different canopy height conditions: (<b>a</b>) R<sup>2</sup>; (<b>b</b>) RMSE.</p>
Full article ">
17 pages, 7929 KiB  
Article
Optimizing Forest Canopy Height Estimation Through Varied Photon-Counting Characteristic Parameter Analysis, Window Size, and Forest Cover
by Jiapeng Huang, Jathun Arachchige Thilini Madushani, Tingting Xia and Xinran Gan
Forests 2024, 15(11), 1957; https://doi.org/10.3390/f15111957 - 7 Nov 2024
Viewed by 970
Abstract
Forests are an important component of the Earth’s ecosystems. Forest canopy height is an important fundamental indicator for quantifying forest ecosystems. The current spaceborne photon-counting Light Detection and Ranging (LiDAR) technique has photon cloud characteristic parameters to estimate forest canopy height, and factors [...] Read more.
Forests are an important component of the Earth’s ecosystems. Forest canopy height is an important fundamental indicator for quantifying forest ecosystems. The current spaceborne photon-counting Light Detection and Ranging (LiDAR) technique has photon cloud characteristic parameters to estimate forest canopy height, and factors such as the sampling window size have not been quantitatively studied. To better understand the precision for estimating canopy height using spaceborne photon-counting LiDAR ICESat-2/ATLAS (Ice, Cloud, and Land Elevation Satellite-2/Advanced Topographic Laser Altimeter System), this study quantified the impact of photon-counting characteristic parameters, sampling window size, and forest cover. Estimation accuracy was evaluated across nine study areas in North America. The findings revealed that when the photon-counting characteristic parameter was set to H70 (70% of canopy height) and the sampling window length was 20 m, the estimation results aligned more closely with the airborne validation data, yielding superior accuracy evaluation indicators with a root mean square error (RMSE) of 4.13 m. Under forest cover of 81%–100%, our algorithms exhibited high estimation accuracy. These study results offer novel perspectives for the application of spaceborne photon-counting LiDAR ICESat-2/ATLAS in forestry. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>A map of the overall locations and ecosystem types.</p>
Full article ">Figure 2
<p>Schematic diagrams of the impact of different photon cloud parameters and windows on estimation accuracy. (<b>a</b>) illustrates the estimation of forest canopy height using different photon cloud characteristic parameters. (<b>b</b>) shows how forest canopy height is estimated using different window sizes.</p>
Full article ">Figure 3
<p>Technical route.</p>
Full article ">Figure 4
<p>The average correlation coefficient (Pearson correlation coefficient, Spearman correlation coefficient, and Kendall correlation coefficient) between two algorithms under different photon cloud characteristic parameter conditions.</p>
Full article ">Figure 5
<p>The average correlation coefficient between two algorithms under different window size conditions.</p>
Full article ">Figure 6
<p>The average correlation coefficient (Pearson correlation coefficient, Spearman correlation coefficient, and Kendall correlation coefficient) between two algorithms under different forest cover.</p>
Full article ">Figure 7
<p>R<sup>2</sup> of different photon-counting characteristic parameters under different forest cover conditions, achieved with this study’s algorithm.</p>
Full article ">Figure 8
<p>RMSEs of different photon-counting characteristic parameters under different forest cover conditions, achieved with this study’s algorithm.</p>
Full article ">Figure 9
<p>R<sup>2</sup> of different photon-counting characteristic parameters under different forest cover conditions, achieved with the NASA method.</p>
Full article ">Figure 10
<p>RMSE of different photon-counting characteristic parameters under different forest cover conditions, achieved with the NASA method.</p>
Full article ">
21 pages, 7085 KiB  
Article
Space-Based Mapping of Pre- and Post-Hurricane Mangrove Canopy Heights Using Machine Learning with Multi-Sensor Observations
by Boya Zhang, Daniel Gann, Shimon Wdowinski, Chaohao Lin, Erin Hestir, Lukas Lamb-Wotton, Khandker S. Ishtiaq, Kaleb Smith and Yuepeng Li
Remote Sens. 2024, 16(21), 3992; https://doi.org/10.3390/rs16213992 - 28 Oct 2024
Viewed by 1067
Abstract
Coastal mangrove forests provide numerous ecosystem services, which can be disrupted by natural disturbances, mainly hurricanes. Canopy height (CH) is a key parameter for estimating carbon storage. Airborne Light Detection and Ranging (LiDAR) is widely viewed as the most accurate method for estimating [...] Read more.
Coastal mangrove forests provide numerous ecosystem services, which can be disrupted by natural disturbances, mainly hurricanes. Canopy height (CH) is a key parameter for estimating carbon storage. Airborne Light Detection and Ranging (LiDAR) is widely viewed as the most accurate method for estimating CH but data are often limited in spatial coverage and are not readily available for rapid impact assessment after hurricane events. Hence, we evaluated the use of systematically acquired space-based Synthetic Aperture Radar (SAR) and optical observations with airborne LiDAR to predict CH across expansive mangrove areas in South Florida that were severely impacted by Category 3 Hurricane Irma in 2017. We used pre- and post-Irma LiDAR-derived canopy height models (CHMs) to train Random Forest regression models that used features of Sentinel-1 SAR time series, Landsat-8 optical, and classified mangrove maps. We evaluated (1) spatial transfer learning to predict regional CH for both time periods and (2) temporal transfer learning coupled with species-specific error correction models to predict post-Irma CH using models trained by pre-Irma data. Model performance of SAR and optical data differed with time period and across height classes. For spatial transfer, SAR data models achieved higher accuracy than optical models for post-Irma, while the opposite was the case for the pre-Irma period. For temporal transfer, SAR models were more accurate for tall trees (>10 m) but optical models were more accurate for short trees. By fusing data of both sensors, spatial and temporal transfer learning achieved the root mean square errors (RMSEs) of 1.9 m and 1.7 m, respectively, for absolute CH. Predicted CH losses were comparable with LiDAR-derived reference values across height and species classes. Spatial and temporal transfer learning techniques applied to readily available spaceborne satellite data can enable conservation managers to assess the impacts of disturbances on regional coastal ecosystems efficiently and within a practical timeframe after a disturbance event. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Florida state boundary and Hurricane Irma track. (<b>b</b>) Level 3 mangrove classification map. (<b>c</b>) Level 4 mangrove species classification. (<b>d</b>) The 30 m G-LiHT footprint of pre-Irma CHM. (<b>e</b>–<b>p</b>) Zoom-in views for four representative sites.</p>
Full article ">Figure 2
<p>Timeline of pre-Irma (orange lines) and post-Irma (purple lines) observations separated by September 2017 Hurricane Irma (the dark vertical line).</p>
Full article ">Figure 3
<p>(<b>a</b>) Flow chart of data filtering. (<b>b</b>) Spatial and temporal transfer learning based on <span class="html-italic">DS2_pre</span> and <span class="html-italic">DS2_post</span> data that are separately used in the spatial transfer but collectively used in the temporal transfer learning. Blue and green rectangles are input variables; white and red rectangles are intermediate products; ovals indicate model processes; pink and orange rectangles are output products.</p>
Full article ">Figure 4
<p>(<b>a</b>) Boxplots for the reference pre- and post-Irma CHM from <span class="html-italic">DS2</span> dataset. Red line indicates the median value; the boxes represent the interquartile range between the first quartile (25th percentile) and the third quartile (75th percentile); the whiskers extend from the edges of the box to the smallest and largest values within 1.5 times the interquartile range. (<b>b</b>) Comparison of backscatter time series and optical observations using a representative pixel from each species. For each subplot, darker color represents pre-Irma values and lighter color post-Irma values. CH values are displayed in the last row.</p>
Full article ">Figure 5
<p>(<b>a</b>) Scatter plot of predicted and referenced CH for pre- and post-Irma from one of the cross-validation evaluation datasets. The yellow lines mark the least-square linear regression model. “BW” in the legend indicates “buttonwood” species. (<b>b</b>) Mean and standard deviation of variable importance of top ten variables using the mixed feature.</p>
Full article ">Figure 6
<p>Scatter plots of prediction error and percentage (Perc) error versus reference CH for both time periods by species. Red lines are the least-squared linear models.</p>
Full article ">Figure 7
<p>Predicted CH (<b>a</b>) pre-Irma, (<b>b</b>) post-Irma, and (<b>c</b>) CH loss (positive values indicate losses). (<b>d</b>) Comparison between mean and standard deviation of predicted and reference CH loss from evaluation datasets in cross-validation. Deep color represents predicted values and light color reference values. Missing data are due to no pixel samples. (<b>e</b>) Local maps of CH loss. White circles in (<b>e3</b>) indicate (<b>left</b>) the bank areas and (<b>right</b>) the boundary between white and red mangroves according to <a href="#remotesensing-16-03992-f001" class="html-fig">Figure 1</a>f.</p>
Full article ">Figure 8
<p>(<b>a</b>) Scatter plot of predicted and corrected CH versus reference post-Irma CH from a cross-validation dataset. (<b>b</b>) Mean and standard deviation of top ten ranking of variable importance using mixed features.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>l</b>) Local maps of post-Irma CH reference, corrected predictions, and errors. White circle outlines indicate areas with large errors. (<b>m</b>) Comparison of the corrected predictions and reference CH losses across pre-Irma canopy height and species classes. Deep color indicates predicted values and light color reference values.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>b</b>) Pre-Irma CH maps for (<b>a</b>) remake of <a href="#remotesensing-16-03992-f007" class="html-fig">Figure 7</a>a and (<b>b</b>) Figure from Jamaluddin et al. (2024) [<a href="#B32-remotesensing-16-03992" class="html-bibr">32</a>]. (<b>c</b>) Remake of CH loss predictions from <a href="#remotesensing-16-03992-f007" class="html-fig">Figure 7</a>c. (<b>d</b>) CH loss predictions from Lagomasino et al. (2021) [<a href="#B10-remotesensing-16-03992" class="html-bibr">10</a>].</p>
Full article ">
18 pages, 16040 KiB  
Article
Unveiling Anomalies in Terrain Elevation Products from Spaceborne Full-Waveform LiDAR over Forested Areas
by Hailan Jiang, Yi Li, Guangjian Yan, Weihua Li, Linyuan Li, Feng Yang, Anxin Ding, Donghui Xie, Xihan Mu, Jing Li, Kaijian Xu, Ping Zhao, Jun Geng and Felix Morsdorf
Forests 2024, 15(10), 1821; https://doi.org/10.3390/f15101821 - 17 Oct 2024
Viewed by 846
Abstract
Anomalies displaying significant deviations between terrain elevation products acquired from spaceborne full-waveform LiDAR and reference elevations are frequently observed in assessment studies. While the predominant focus is on “normal” data, recognizing anomalies within datasets obtained from the Geoscience Laser Altimeter System (GLAS) and [...] Read more.
Anomalies displaying significant deviations between terrain elevation products acquired from spaceborne full-waveform LiDAR and reference elevations are frequently observed in assessment studies. While the predominant focus is on “normal” data, recognizing anomalies within datasets obtained from the Geoscience Laser Altimeter System (GLAS) and the Global Ecosystem Dynamics Investigation (GEDI) is essential for a comprehensive understanding of widely used spaceborne full-waveform data, which not only facilitates optimal data utilization but also enhances the exploration of potential applications. Nevertheless, our comprehension of anomalies remains limited as they have received scant specific attention. Diverging from prevalent practices of directly eliminating outliers, we conducted a targeted exploration of anomalies in forested areas using both transmitted and return waveforms from the GLAS and the GEDI in conjunction with airborne LiDAR point cloud data. We unveiled that elevation anomalies stem not from the transmitted pulses or product algorithms, but rather from scattering sources. We further observed similarities between the GLAS and the GEDI despite their considerable disparities in sensor parameters, with the waveforms characterized by a low signal-to-noise ratio and a near exponential decay in return energy; specifically, return signals of anomalies originated from clouds rather than the land surface. This discovery underscores the potential of deriving cloud-top height from spaceborne full-waveform LiDAR missions, particularly the GEDI, suggesting promising prospects for applying GEDI data in atmospheric science—an area that has received scant attention thus far. To mitigate the impact of abnormal return waveforms on diverse land surface studies, we strongly recommend incorporating spaceborne LiDAR-offered terrain elevation in data filtering by establishing an elevation-difference threshold against a reference elevation. This is especially vital for studies concerning forest parameters due to potential cloud interference, yet a consensus has not been reached within the community. Full article
Show Figures

Figure 1

Figure 1
<p>Study area and the geolocation of GLAS and GEDI data.</p>
Full article ">Figure 2
<p>Flowchart of this study.</p>
Full article ">Figure 3
<p>Spatial distribution of terrain elevation anomalies in the GLAS and GEDI datasets.</p>
Full article ">Figure 4
<p>Scatter plot of terrain elevation estimates obtained from GLAS (<b>a</b>) and GEDI (<b>b</b>) vs. the terrain elevation derived from airborne laser scanning (ALS) as a reference. A0 denotes the default algorithm of the GEDI L2A product.</p>
Full article ">Figure 5
<p>Details of terrain elevation outliers from GLAS: scatter plot of terrain elevation from data acquired during nighttime (<b>a</b>) and daytime (<b>b</b>) before removing outliers, scatter plot (<b>c</b>), transmitted waveforms (<b>d</b>), the histogram of the data acquisition time (<b>e</b>), and the histogram of Signal-to-Noise Ratio (SNR) (<b>f</b>) of source laser shot of the outliers.</p>
Full article ">Figure 6
<p>Details of terrain elevation outliers from GEDI: scatter plot of terrain elevation from data acquired during nighttime (<b>a</b>) and daytime (<b>b</b>) before removing outliers, scatter plot (<b>c</b>), transmitted waveforms (<b>d</b>), the histogram (<b>e</b>) of the beam type (<b>e1</b>) and data acquisition time (<b>e2</b>), and the histogram of sensitivity (<b>f</b>) of source laser shot of the outliers.</p>
Full article ">Figure 7
<p>Examples with small (upper panel) and large (lower panel) terrain elevation error: the three-dimensional scene (<b>left</b>), the transmitted waveform (<b>middle</b>), and the return waveform (<b>right</b>) of GLAS and GEDI with the terrain elevation from product and airborne laser scanning (ALS) data illustrated. In the lower panel (<b>right</b>), the ALS terrain elevation is not indicated since the GLAS or GEDI terrain elevation exceeds ALS by more than 330 m (see outlier-27 and outlier-12 indicated in green circles in <a href="#forests-15-01821-f005" class="html-fig">Figure 5</a> and <a href="#forests-15-01821-f006" class="html-fig">Figure 6</a>c). A&lt;<span class="html-italic">n</span>&gt; (<span class="html-italic">n</span>: 1–6) denotes the terrain elevation from six different algorithm groups of GEDI.</p>
Full article ">Figure 8
<p>Scatter plot of canopy height estimates for laser shots of terrain elevation anomalies obtained from GEDI L2A product versus the canopy height derived from airborne laser scanning (ALS) as a reference (the legend of the point density applies to all the figures). A0 denotes the default algorithm (<b>a</b>), and A&lt;<span class="html-italic">n</span>&gt; (<span class="html-italic">n</span>: 1–6) denotes the other six algorithm groups (<b>b</b>–<b>g</b>).</p>
Full article ">Figure A1
<p>Probability density of “sensitivity” of “power” and “coverage” beams estimated by different algorithms (<b>a</b>–<b>f</b>) of GEDI using the data with “sensitivity &gt; 0.90” in all footprints. A0 denotes the default algorithm setting (<b>a</b>), and A&lt;n&gt; (n: 1–6) denotes the other six algorithm groups (<b>b</b>–<b>g</b>).</p>
Full article ">Figure A2
<p>Original GLAS (<b>upper panel</b>) and GEDI (<b>lower panel</b>) waveform examples of terrain elevation anomalies with terrain elevation provided by GLAS and GEDI product indicated. A0 denotes the default algorithm, and A&lt;<span class="html-italic">n</span>&gt; (<span class="html-italic">n</span>: 1–6) denotes the other six algorithm groups of GEDI.</p>
Full article ">
19 pages, 3356 KiB  
Article
The First Validation of Aerosol Optical Parameters Retrieved from the Terrestrial Ecosystem Carbon Inventory Satellite (TECIS) and Its Application
by Yijie Ren, Binglong Chen, Lingbing Bu, Gen Hu, Jingyi Fang and Pasindu Liyanage
Remote Sens. 2024, 16(19), 3689; https://doi.org/10.3390/rs16193689 - 3 Oct 2024
Viewed by 569
Abstract
In August 2022, China successfully launched the Terrestrial Ecosystem Carbon Inventory Satellite (TECIS). The primary payload of this satellite is an onboard multi-beam lidar system, which is capable of observing aerosol optical parameters on a global scale. This pioneering study used the Fernald [...] Read more.
In August 2022, China successfully launched the Terrestrial Ecosystem Carbon Inventory Satellite (TECIS). The primary payload of this satellite is an onboard multi-beam lidar system, which is capable of observing aerosol optical parameters on a global scale. This pioneering study used the Fernald forward integration method to retrieve aerosol optical parameters based on the Level 2 data of the TECIS, including the aerosol depolarization ratio, aerosol backscatter coefficient, aerosol extinction coefficient, and aerosol optical depth (AOD). The validation of the TECIS-retrieved aerosol optical parameters was conducted using CALIPSO Level 1 and Level 2 data, with relative errors within 30%. A comparison of the AOD retrieved from the TECIS with the AERONET and MODIS AOD products yielded correlation coefficients greater than 0.7 and 0.6, respectively. The relative error of aerosol optical parameter profiles compared with ground-based measurements for CALIPSO was within 40%. Additionally, the correlation coefficients R2 with MODIS and AERONET AOD were approximately between 0.5 and 0.7, indicating the high accuracy of TECIS retrievals. Utilizing the TECIS retrieval results, combined with ground air quality monitoring data and HYSPLIT outcomes, a typical dust transport event was analyzed from 2 to 7 April 2023. The results indicate that dust was transported from the Taklamakan Desert in Xinjiang, China, to Henan and Anhui provinces, with a gradual decrease in the aerosol depolarization ratio and backscatter coefficient during the transport process, causing varying degrees of pollution in the downstream regions. This research verifies the accuracy of the retrieval algorithm through multi-source data comparison and demonstrates the potential application of the TECIS in the field of aerosol science for the first time. It enables the fine-scale regional monitoring of atmospheric aerosols and provides reliable data support for the three-dimensional distribution of global aerosols and related scientific applications. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>A flowchart of the TECIS retrieval algorithm.</p>
Full article ">Figure 2
<p>Trajectory of CALIPSO and TECIS.</p>
Full article ">Figure 3
<p>Total attenuated backscatter coefficient obtained from TECIS and CALIPSO. (<b>a</b>) TECIS, (<b>b</b>) CALIPSO.</p>
Full article ">Figure 4
<p>SNR of total attenuated backscatter coefficient obtained from TECIS and CALIPSO. (<b>a</b>) TECIS, (<b>b</b>) CALIPSO.</p>
Full article ">Figure 5
<p>A comparison of total attenuation backscatter coefficient mean profiles between the TECIS and CALIPSO at 13° to 14°N; the shaded area represents the standard deviation of the two satellites. The blue solid line represents the TECIS result, and the red solid line represents the CALIPSO result.</p>
Full article ">Figure 6
<p>(<b>a</b>) Profile of aerosol depolarization ratio of TECIS, (<b>b</b>) profile of aerosol depolarization ratio of CALIPSO, (<b>c</b>) profile of aerosol backscatter coefficient of TECIS, (<b>d</b>) profile of aerosol backscatter coefficient of CALIPSO, (<b>e</b>) profile of aerosol extinction coefficient of TECIS, (<b>f</b>) profile of aerosol extinction coefficient of CALIPSO.</p>
Full article ">Figure 7
<p>A comparison of aerosol optical parameter mean profiles between the TECIS and CALIPSO at 13° to 14°N, where the blue solid line represents the TECIS result, the red solid line represents the CALIPSO result, and the shaded area represents the standard deviation within the average range of the two satellites. (<b>a</b>) Aerosol depolarization ratio, (<b>b</b>) aerosol backscatter coefficient, (<b>c</b>) aerosol extinction coefficient.</p>
Full article ">Figure 8
<p>Relative error of retrieval results between TECIS and CALIPSO.</p>
Full article ">Figure 9
<p>TECIS 532 nm AOD retrievals against AERONET AOD during April to June 2023; the dashed line is the linear fit described by the regression equation; the black line is the 1:1 line.</p>
Full article ">Figure 10
<p>A scatterplot comparison of TECIS AOD data against MODIS AOD data during April to June 2023; the color scale represents the fraction of the total data. (<b>a</b>) North Africa, (<b>b</b>) the Middle East, (<b>c</b>) North America, (<b>d</b>) Central Asia.</p>
Full article ">Figure 11
<p>TECIS 1064 nm total attenuation backscattering coefficient and HYSPLIT backward tracking from 2 to 7 April 2023 (blue, red, and black represent backward tracking at heights of 3 km, 2 km, and 1 km, respectively).</p>
Full article ">Figure 12
<p>Variations in PM10 and PM2.5 concentrations from 2 to 7 April 2023. (<b>a</b>) PM10, (<b>b</b>) PM2.5.</p>
Full article ">Figure 13
<p>Optical parameters obtained by TECIS inversion from 2 to 7 April 2023. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) show backscattering coefficient; (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) show depolarization ratio.</p>
Full article ">Figure 13 Cont.
<p>Optical parameters obtained by TECIS inversion from 2 to 7 April 2023. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>) show backscattering coefficient; (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) show depolarization ratio.</p>
Full article ">
24 pages, 27095 KiB  
Article
Examining the Impact of Topography and Vegetation on Existing Forest Canopy Height Products from ICESat-2 ATLAS/GEDI Data
by Yisa Li, Dengsheng Lu, Yagang Lu and Guiying Li
Remote Sens. 2024, 16(19), 3650; https://doi.org/10.3390/rs16193650 - 30 Sep 2024
Viewed by 959
Abstract
Forest canopy height (FCH) is an important variable for estimating forest biomass and ecosystem carbon sequestration. Spaceborne LiDAR data have been used to create wall-to-wall FCH maps, such as the forest tree height map of China (FCHChina), Global Forest Canopy Height 2020 (GFCH2020), [...] Read more.
Forest canopy height (FCH) is an important variable for estimating forest biomass and ecosystem carbon sequestration. Spaceborne LiDAR data have been used to create wall-to-wall FCH maps, such as the forest tree height map of China (FCHChina), Global Forest Canopy Height 2020 (GFCH2020), and Global Forest Canopy Height 2019 (GFCH2019). However, these products lack comprehensive assessment. This study used airborne LiDAR data from various topographies (e.g., plain, hill, and mountain) to assess the impacts of different topographical and vegetation characteristics on spaceborne LiDAR-derived FCH products. The results show that GEDI–FCH demonstrates better accuracy in plain and hill regions, while ICESat-2 ATLAS–FCH shows superior accuracy in the mountainous region. The difficulty in accurately capturing photons from sparse tree canopies by ATLAS and the geolocation errors of GEDI has led to partial underestimations of FCH products in plain areas. Spaceborne LiDAR FCH retrievals are more accurate in hilly regions, with a root mean square error (RMSE) of 4.99 m for ATLAS and 3.85 m for GEDI. GEDI–FCH is significantly affected by slope in mountainous regions, with an RMSE of 13.26 m. For wall-to-wall FCH products, the availability of FCH data is limited in plain areas. Optimal accuracy is achieved in hilly regions by FCHChina, GFCH2020, and GFCH2019, with RMSEs of 5.52 m, 5.07 m, and 4.85 m, respectively. In mountainous regions, the accuracy of wall-to-wall FCH products is influenced by factors such as tree canopy coverage, forest cover types, and slope. However, some of these errors may stem from directly using current ATL08 and GEDI L2A FCH products for mountainous FCH estimation. Introducing accurate digital elevation model (DEM) data can improve FCH retrieval from spaceborne LiDAR to some extent. This research improves our understanding of the existing FCH products and provides valuable insights into methods for more effectively extracting accurate FCH from spaceborne LiDAR data. Further research should focus on developing suitable approaches to enhance the FCH retrieval accuracy from spaceborne LiDAR data and integrating multi-source data and modeling algorithms to produce accurate wall-to-wall FCH distribution in a large area. Full article
(This article belongs to the Special Issue Lidar for Forest Parameters Retrieval)
Show Figures

Figure 1

Figure 1
<p>The selected three study areas in Anhui Province: (<b>a</b>) Landform types; (<b>b</b>) Lixin County in plain region; (<b>c</b>) Lujiang County in hilly region; (<b>d</b>) Huangshan District in mountainous region, overlaid with ATLAS and GEDI tracks.</p>
Full article ">Figure 2
<p>Profiles of spaceborne LiDAR data and corresponding ALS data along the tracks in mountainous region (<b>a</b>) ICESat-2 ATLAS, (<b>b</b>) GEDI.</p>
Full article ">Figure 3
<p>Comparison of ATLAS photons and ALS DSM and DEM among plain, hill, and mountain with different slopes and canopy cover levels. (<b>a1</b>–<b>a4</b>) different canopy levels in plain region; (<b>b1</b>–<b>b4</b>) different combinations of slopes and canopy covers in hilly region; (<b>c1</b>–<b>c4</b>) different combinations of slopes and canopy covers in mountainous region.</p>
Full article ">Figure 4
<p>Comparison of GEDI elevation and ALS DSM and DEM among plain, hill and mountain with different slopes and canopy cover levels. (<b>a1</b>–<b>a4</b>) different canopy levels in plain region; (<b>b1</b>–<b>b4</b>) different combinations of slopes and canopy covers in hilly region; (<b>c1</b>–<b>c4</b>) different combinations of slopes and canopy covers in mountainous region.</p>
Full article ">Figure 5
<p>Comparison of spaceborne LiDAR FCH and corresponding ALS FCH (<b>a</b> and <b>b</b> represent ATLAS FCH and GEDI FCH; <b>1</b>, <b>2,</b> and <b>3</b> represent plain, hill, and mountain regions).</p>
Full article ">Figure 6
<p>Root mean square error (RMSE), relative RMSE (rRMSE), and bias of ATLAS forest canopy height (FCH) and GEDI FCH under different terrain slope levels and canopy cover levels in mountain region (Huangshan District) (for canopy cover, only footprints with slopes &lt; 20° were used for analysis) (<b>a</b>, <b>b</b>, and <b>c</b> represent RMSE, rRMSE, and bias; <b>1</b> and <b>2</b> represent slope and canopy cover).</p>
Full article ">Figure 7
<p>Comparison of three wall-to-wall FCH products (<b>a</b>—FCHChina; <b>b</b>—GFCH2020; and <b>c</b>—GFCH2019) with corresponding ALS FCH based on different topographies (<b>1</b>, <b>2</b>, and <b>3</b> represent plain, hill, and mountain regions).</p>
Full article ">Figure 8
<p>Residuals of three FCH products ((<b>a</b>) FCHChina, (<b>b</b>) GFCH2020, and (<b>c</b>) GFCH2019) at different slope intervals (Residual here is FCH maps—ALS referenced values).</p>
Full article ">Figure 9
<p>Comparison of wall-to-wall FCH products with ALS FCH in different slope intervals (<b>a</b>, <b>b</b> and <b>c</b> represent FCHChina, GFCH2020, and GFCH2019; <b>1</b>, <b>2</b>, <b>3</b>, <b>4</b>, <b>5</b>, and <b>6</b> represent five slope ranges: 0°–10°, 10°–20°, 20°–30°, 30°–40°, 40°–50°, and 50°–60°).</p>
Full article ">Figure 10
<p>Assessment of wall-to-wall FCH products with ALS FCH under different forest types (<b>a</b>, <b>b</b>, and <b>c</b> represent FCHChina, GFCH2020, and GFCH2019; <b>1</b>, <b>2</b>, <b>3</b>, and <b>4</b> represent Chinese fir, Masson pine, Moso bamboo, and broadleaf forests).</p>
Full article ">Figure 11
<p>Assessment of wall-to-wall FCH products with ALS FCH under different canopy covers (<b>a</b>, <b>b</b>, and <b>c</b> represent FCHChina, GFCH2020, and GFCH2019; <b>1</b>, <b>2</b>, and <b>3</b> represent low, medium, and high canopy covers).</p>
Full article ">Figure 12
<p>Comparison of original FCH and modified one using ALS data, (<b>a</b>) original ATLAS FCH, (<b>b</b>) improved ATLAS FCH, (<b>c</b>) original GEDI FCH, (<b>d</b>) improved GEDI FCH).</p>
Full article ">Figure 13
<p>Residual distribution of ATLAS FCH and ALS FCH under different canopy photon numbers in plain area.</p>
Full article ">
19 pages, 15677 KiB  
Article
Automatic Correction of Time-Varying Orbit Errors for Single-Baseline Single-Polarization InSAR Data Based on Block Adjustment Model
by Huacan Hu, Haiqiang Fu, Jianjun Zhu, Zhiwei Liu, Kefu Wu, Dong Zeng, Afang Wan and Feng Wang
Remote Sens. 2024, 16(19), 3578; https://doi.org/10.3390/rs16193578 - 26 Sep 2024
Viewed by 841
Abstract
Orbit error is one of the primary error sources of interferometric synthetic aperture radar (InSAR) and differential InSAR (D-InSAR) measurements, arising from inaccurate orbit determination of SAR platforms. Typically, orbit error in the interferogram can be estimated using polynomial models. However, correcting for [...] Read more.
Orbit error is one of the primary error sources of interferometric synthetic aperture radar (InSAR) and differential InSAR (D-InSAR) measurements, arising from inaccurate orbit determination of SAR platforms. Typically, orbit error in the interferogram can be estimated using polynomial models. However, correcting for orbit errors with significant time-varying characteristics presents two main challenges: (1) the complexity and variability of the azimuth time-varying orbit errors make it difficult to accurately model them using a set of polynomial coefficients; (2) existing patch-based polynomial models rely on empirical segmentation and overlook the time-varying characteristics, resulting in residual orbital error phase. To overcome these problems, this study proposes an automated block adjustment framework for estimating time-varying orbit errors, incorporating the following innovations: (1) the differential interferogram is divided into several blocks along the azimuth direction to model orbit error separately; (2) automated segmentation is achieved by extracting morphological features (i.e., peaks and troughs) from the azimuthal profile; (3) a block adjustment method combining control points and connection points is proposed to determine the model coefficients of each block for the orbital error phase estimation. The feasibility of the proposed method was verified by repeat-pass L-band spaceborne and P-band airborne InSAR data, and finally, the InSAR digital elevation model (DEM) was generated for performance evaluation. Compared with the high-precision light detection and ranging (LiDAR) elevation, the root mean square error (RMSE) of InSAR DEM was reduced from 18.27 m to 7.04 m in the spaceborne dataset and from 7.83~14.97 m to 3.36~6.02 m in the airborne dataset. Then, further analysis demonstrated that the proposed method outperforms existing algorithms under single-baseline and single-polarization conditions. Moreover, the proposed method is applicable to both spaceborne and airborne InSAR data, demonstrating strong versatility and potential for broader applications. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of the proposed method for estimating the time-varying orbital error phase.</p>
Full article ">Figure 2
<p>Schematic diagram of differential interferogram division by the proposed method. (<b>a</b>) is the differential interferogram, (<b>b</b>) is the profile of the time-varying orbital error phase along the azimuth at the dotted line shown in (<b>a</b>), (<b>c</b>) is a schematic diagram of different blocks, and (<b>d</b>) is the overlap area between different blocks and the distribution of control points and connection points.</p>
Full article ">Figure 3
<p>Geographical location of the test sites: (<b>a</b>) Hunan and (<b>b</b>) Krycklan. The red boxes represent LuTan-1 InSAR data and E-SAR data, the blue box represents airborne LiDAR data, and the yellow dots represent the footprint of ICESat-2 ATL08 elevation.</p>
Full article ">Figure 4
<p>Orbital error phase analysis of LT-1 InSAR data from the Hunan test site: (<b>a</b>) interferometric coherence, (<b>b</b>) differential interferometric phase, (<b>c</b>,<b>d</b>) are the profiles of the orbital error phase in the azimuth and range directions, respectively.</p>
Full article ">Figure 5
<p>Time-varying orbit error phase estimation results: (<b>a</b>) original differential interferometric phase, (<b>b</b>) estimated orbital error phase, (<b>c</b>) differential interferometric phase after removing orbit error.</p>
Full article ">Figure 6
<p>DEM elevation validation. (<b>a</b>) InSAR DEM from interferogram inversion after removing orbit error, (<b>b</b>) difference between InSAR DEM and external DEM, (<b>c</b>) error histogram of InSAR DEM relative to ICESat−2 elevation.</p>
Full article ">Figure 7
<p>Analysis of airborne P−band time−varying orbit errors in the Krycklan test site. (<b>a</b>) Airborne SAR intensity map, (<b>b</b>) azimuth profiles of five differential interferograms, (<b>c</b>) azimuth profile of interferometric pair 0101–0103 and extracted peaks and troughs, (<b>d</b>) range profile of interferometric pair 0101–0103.</p>
Full article ">Figure 8
<p>Results of the proposed method for estimating airborne time–varying orbital error phase. (<b>a1</b>–<b>a5</b>) are original differential interferometric phases, (<b>b1</b>–<b>b5</b>) are the estimated orbital error phases, (<b>c1</b>–<b>c5</b>) are differential interferometric phases after removing orbit error phases. From left to right, the five interferometric pairs shown in <a href="#remotesensing-16-03578-t001" class="html-table">Table 1</a> are represented.</p>
Full article ">Figure 9
<p>(<b>a</b>) InSAR DEM estimated by the interferometric pair numbered 0103–0111, (<b>b</b>–<b>f</b>) differences between InSAR DEM and LiDAR DTM estimated after correcting orbit errors for the five interferometric pairs in <a href="#remotesensing-16-03578-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>e</b>) Error statistical histograms of InSAR DEM and LiDAR DTM estimated before and after orbital error correction for the five interferometric pairs in <a href="#remotesensing-16-03578-t001" class="html-table">Table 1</a>.</p>
Full article ">
15 pages, 6660 KiB  
Article
Forest Canopy Height Estimation Combining Dual-Polarization PolSAR and Spaceborne LiDAR Data
by Yao Tong, Zhiwei Liu, Haiqiang Fu, Jianjun Zhu, Rong Zhao, Yanzhou Xie, Huacan Hu, Nan Li and Shujuan Fu
Forests 2024, 15(9), 1654; https://doi.org/10.3390/f15091654 - 19 Sep 2024
Viewed by 963
Abstract
Forest canopy height data are fundamental parameters of forest structure and are critical for understanding terrestrial carbon stock, global carbon cycle dynamics and forest productivity. To address the limitations of retrieving forest canopy height using conventional PolInSAR-based methods, we proposed a method to [...] Read more.
Forest canopy height data are fundamental parameters of forest structure and are critical for understanding terrestrial carbon stock, global carbon cycle dynamics and forest productivity. To address the limitations of retrieving forest canopy height using conventional PolInSAR-based methods, we proposed a method to estimate forest height by combining single-temporal polarimetric synthetic aperture radar (PolSAR) images with sparse spaceborne LiDAR (forest height) measurements. The core idea of our method is that volume scattering energy variations which are linked to forest canopy height occur during radar acquisition. Specifically, our methodology begins by employing a semi-empirical inversion model directly derived from the random volume over ground (RVoG) formulation to establish the relationship between forest canopy height, volume scattering energy and wave extinction. Subsequently, PolSAR decomposition techniques are used to extract canopy volume scattering energy. Additionally, machine learning is employed to generate a spatially continuous extinction coefficient product, utilizing sparse LiDAR samples for assistance. Finally, with the derived inversion model and the resulting model parameters (i.e., volume scattering power and extinction coefficient), forest canopy height can be estimated. The performance of the proposed forest height inversion method is illustrated with L-band NASA/JPL UAVSAR from AfriSAR data conducted over the Gabon Lope National Park and airborne LiDAR data. Compared to high-accuracy airborne LiDAR data, the obtained forest canopy height from the proposed approach exhibited higher accuracy (R2 = 0.92, RMSE = 6.09 m). The results demonstrate the potential and merit of the synergistic combination of PolSAR (volume scattering power) and sparse LiDAR (forest height) measurements for forest height estimation. Additionally, our approach achieves good performance in forest height estimation, with accuracy comparable to that of the multi-baseline PolInSAR-based inversion method (RMSE = 5.80 m), surpassing traditional PolSAR-based methods with an accuracy of 10.86 m. Given the simplicity and efficiency of the proposed method, it has the potential for large-scale forest height estimation applications when only single-temporal dual-polarization acquisitions are available. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>A flowchart of the methodology proposed for the estimation of forest canopy height.</p>
Full article ">Figure 2
<p>The geolocation of the study area: (<b>a</b>) optical imagery; (<b>b</b>) the digital elevation model (DEM) of the study area. The orange rectangles in (<b>a</b>,<b>b</b>) indicate the coverage range of these airborne PolSAR data.</p>
Full article ">Figure 3
<p>Datasets: (a) multi-looked and geocoded SAR image in Pauli basis color combination; (<b>b</b>) ICESat-2 ATL08 sampling points; (<b>c</b>) LVIS forest height.</p>
Full article ">Figure 4
<p>(<b>a</b>) Volume scattering power; (<b>b</b>) extinction coefficient.</p>
Full article ">Figure 5
<p>Importance ranking of each variable in the extinction coefficient estimation model.</p>
Full article ">Figure 6
<p>(<b>a</b>) Forest height map derived by proposed method; (<b>b</b>) validation plots of the forest height inversion, where the color transition from blue to red indicates an increase density of points.</p>
Full article ">Figure 7
<p>(<b>a</b>) Forest height derived via PolSAR inversion method in [<a href="#B18-forests-15-01654" class="html-bibr">18</a>], and (<b>b</b>) scatterplot of validation results.</p>
Full article ">
23 pages, 11057 KiB  
Article
Denoising of Photon-Counting LiDAR Bathymetry Based on Adaptive Variable OPTICS Model and Its Accuracy Assessment
by Peize Li, Yangrui Xu, Yanpeng Zhao, Kun Liang and Yuanjie Si
Remote Sens. 2024, 16(18), 3438; https://doi.org/10.3390/rs16183438 - 16 Sep 2024
Viewed by 762
Abstract
Spaceborne photon-counting LiDAR holds significant potential for shallow-water bathymetry. However, the received photon data often contain substantial noise, complicating the extraction of elevation information. Currently, a denoising algorithm named ordering points to identify the clustering structure (OPTICS) draws people’s attention because of its [...] Read more.
Spaceborne photon-counting LiDAR holds significant potential for shallow-water bathymetry. However, the received photon data often contain substantial noise, complicating the extraction of elevation information. Currently, a denoising algorithm named ordering points to identify the clustering structure (OPTICS) draws people’s attention because of its strong performance under high background noise. However, this algorithm’s fixed input variables can lead to inaccurate photon distribution parameters in areas near the water bottom, which results in inadequate denoising in these areas, affecting bathymetric accuracy. To address this issue, an Adaptive Variable OPTICS (AV-OPTICS) model is proposed in this paper. Unlike the traditional OPTICS model with fixed input variables, the proposed model dynamically adjusts input variables based on point cloud distribution. This adjustment ensures accurate measurement of photon distribution parameters near the water bottom, thereby enhancing denoising effects in these areas and improving bathymetric accuracy. The findings indicate that, compared to traditional OPTICS methods, AV-OPTICS achieves higher F1-values and lower cohesions, demonstrating better denoising performance near the water bottom. Furthermore, this method achieves an average MAE of 0.28 m and RMSE of 0.31 m, indicating better bathymetric accuracy than traditional OPTICS methods. This study provides a promising solution for shallow-water bathymetry based on photon-counting LiDAR data. Full article
(This article belongs to the Special Issue Satellite Remote Sensing for Ocean and Coastal Environment Monitoring)
Show Figures

Figure 1

Figure 1
<p>Study area and the detection tracks of ATLAS represented by six different colors of dashed lines.</p>
Full article ">Figure 2
<p>Water depth results of study area for ALB reference data.</p>
Full article ">Figure 3
<p>The elevation distribution histogram of ATL03 photon data.</p>
Full article ">Figure 4
<p>The contours of the water bottom terrain under different scenarios. (<b>a</b>) relatively flat water bottom terrains; (<b>b</b>) complex water bottom terrains.</p>
Full article ">Figure 5
<p>The distance of point <span class="html-italic">o</span> and <span class="html-italic">w</span> under the definition of OPTICS algorithm.</p>
Full article ">Figure 6
<p>The spatial geometric relationships of refraction correction under different slope angles <math display="inline"><semantics> <mrow> <mi>φ</mi> </mrow> </semantics></math>. The green and red vectors correspond to the original coordinate and corrected coordinate of water bottom photons, respectively [<a href="#B38-remotesensing-16-03438" class="html-bibr">38</a>].</p>
Full article ">Figure 7
<p>Denoising effects of our method and traditional OPTICS in different scenes. (<b>a</b>) 20190119gt3l—Raw data; (<b>b</b>) 20190119gt3l—Our method; (<b>c</b>) 20190119gt3l—Traditional OPTICS; (<b>d</b>) 20181024gt3r—Raw data; (<b>e</b>) 20181024gt3r—Our method; (<b>f</b>) 20181024gt3r—Traditional OPTICS; (<b>g</b>) 20200717gt3l—Raw data; (<b>h</b>) 20200717gt3l—Our method; (<b>i</b>) 20200717gt3l—Traditional OPTICS.</p>
Full article ">Figure 8
<p>Comparison of denoising details of the two methods in different scenarios. (<b>a</b>) 20181024gt3r—Our method; (<b>b</b>) 20181024gt3r—Traditional OPTICS; (<b>c</b>) 20190420gt2l—Our method; (<b>d</b>) 20190420gt2l—Traditional OPTICS.</p>
Full article ">Figure 8 Cont.
<p>Comparison of denoising details of the two methods in different scenarios. (<b>a</b>) 20181024gt3r—Our method; (<b>b</b>) 20181024gt3r—Traditional OPTICS; (<b>c</b>) 20190420gt2l—Our method; (<b>d</b>) 20190420gt2l—Traditional OPTICS.</p>
Full article ">Figure 9
<p>Coordinate correction and fitting profiles of the signal photons. (<b>a</b>) 20190119gt3l—Coordinate correction; (<b>b</b>) 20190119gt3l—Fitting profiles; (<b>c</b>) 20181024gt3r—Coordinate correction; (<b>d</b>) 20181024gt3r—Fitting profiles.</p>
Full article ">Figure 9 Cont.
<p>Coordinate correction and fitting profiles of the signal photons. (<b>a</b>) 20190119gt3l—Coordinate correction; (<b>b</b>) 20190119gt3l—Fitting profiles; (<b>c</b>) 20181024gt3r—Coordinate correction; (<b>d</b>) 20181024gt3r—Fitting profiles.</p>
Full article ">Figure 10
<p>Bathymetric accuracy validation and comparison of our method and traditional OPTICS. (<b>a</b>,<b>b</b>) on the first row correspond to 20190119gt3l, while (<b>c</b>,<b>d</b>) on the second row correspond to 20190420gt2l. The red and black points represent the bathymetric results of the corresponding method and in situ data, respectively. (<b>a</b>) 20190119gt3l—Our method; (<b>b</b>) 20190119gt3l—Traditional OPTICS; (<b>c</b>) 20190420gt2l—Our method; (<b>d</b>) 20190420gt2l—Traditional OPTICS.</p>
Full article ">Figure 11
<p>Idealized model of elliptic filter in vertical direction. The water bottom contour within the black block is regarded as a gray rectangle. The yellow, red, blue, and green lines are the idealized elliptical filters with different lengths of semi-minor axis, respectively.</p>
Full article ">Figure 12
<p>Idealized model of elliptic filter in horizontal direction. The water bottom contour within the black block is regarded as a gray rectangle. The yellow and red lines are the idealized elliptical filters with different lengths of semi-major axis, respectively.</p>
Full article ">Figure 13
<p>Deviation percentage between ICESat-2 results and ALB in situ data, with an interval of 0.5 m for each histogram column. (<b>a</b>,<b>b</b>) on the first row correspond to 20190119gt3l, while (<b>c</b>,<b>d</b>) on the second row correspond to 20201016gt2l. (<b>a</b>) 20190119gt3l—Our method; (<b>b</b>) 20190119gt3l—Traditional OPTICS; (<b>c</b>) 20201016gt2l—Our method; (<b>d</b>) 20201016gt2l—Traditional OPTICS.</p>
Full article ">Figure 14
<p>Bathymetric accuracy comparison of our method and that without coordinate correction. (<b>a</b>–<b>c</b>) on the first row correspond to 20190119gt3l, while (<b>d</b>–<b>f</b>) on the second row correspond to 20181024gt3r. The red and black lines represent the bathymetric results of the corresponding method and in situ data, respectively. (<b>a</b>) 20190119gt3r—Our method; (<b>b</b>) Without refraction correction; (<b>c</b>) Without tidal correction; (<b>d</b>) 20181024gt3r—Our method; (<b>e</b>) Without refraction correction; (<b>f</b>) Without tide correction.</p>
Full article ">
22 pages, 4955 KiB  
Article
Statistically Resolved Planetary Boundary Layer Height Diurnal Variability Using Spaceborne Lidar Data
by Natalia Roldán-Henao, John E. Yorks, Tianning Su, Patrick A. Selmer and Zhanqing Li
Remote Sens. 2024, 16(17), 3252; https://doi.org/10.3390/rs16173252 - 2 Sep 2024
Viewed by 1023
Abstract
The Planetary Boundary Layer Height (PBLH) significantly impacts weather, climate, and air quality. Understanding the global diurnal variation of the PBLH is particularly challenging due to the necessity of extensive observations and suitable retrieval algorithms that can adapt to diverse thermodynamic and dynamic [...] Read more.
The Planetary Boundary Layer Height (PBLH) significantly impacts weather, climate, and air quality. Understanding the global diurnal variation of the PBLH is particularly challenging due to the necessity of extensive observations and suitable retrieval algorithms that can adapt to diverse thermodynamic and dynamic conditions. This study utilized data from the Cloud-Aerosol Transport System (CATS) to analyze the diurnal variation of PBLH in both continental and marine regions. By leveraging CATS data and a modified version of the Different Thermo-Dynamics Stability (DTDS) algorithm, along with machine learning denoising, the study determined the diurnal variation of the PBLH in continental mid-latitude and marine regions. The CATS DTDS-PBLH closely matches ground-based lidar and radiosonde measurements at the continental sites, with correlation coefficients above 0.6 and well-aligned diurnal variability, although slightly overestimated at nighttime. In contrast, PBLH at the marine site was consistently overestimated due to the viewing geometry of CATS and complex cloud structures. The study emphasizes the importance of integrating meteorological data with lidar signals for accurate and robust PBLH estimations, which are essential for effective boundary layer assessment from satellite observations. Full article
(This article belongs to the Special Issue Observation of Atmospheric Boundary-Layer Based on Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The locations of the two primary ARM sites used in this study. The red and magenta dots correspond to the Southern Great Plains (SGP) and Eastern North Atlantic (ENA) central locations, respectively. The squared boxes enclosing the dots correspond to the geographical region where CATS backscatter data is taken.</p>
Full article ">Figure 2
<p>(<b>a</b>) CATS Total Attenuated Backscatter and (<b>b</b>) CATS Denoised photon counts on 11 July 2015.</p>
Full article ">Figure 3
<p>Schematic diagram of the modified DTDS algorithm for the space-borne lidars. In this diagram, H(i) represents the Planetary Boundary Layer Height (PBLH) at each spatial interval i; CBH indicates the cloud base in single-layer clouds with a thickness smaller than 1 km. The PBLH selection identify the local maximum positions (LMPs) based on the spatial continuity. PBLH under the cloudy condition is determined by assessing cloud-surface coupling and utilizing information on the adjacent (nearby pixels) PBLH, cloud base height (CBH) and the lifted condensation level (LCL). Empirical parameters A1, A2, and A5 are set to 0.7, 0.2, and 1.1, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Example of the computed PBLH on 10 January 2016, the background colors correspond to the total attenuated backscatter measured by CATS over the Southern Great Plains (SGP). Black dots correspond to the Different Thermo-Dynamics Stability (DTDS) Planetary Boundary Layer Height (DTDS-PBLH), yellow dots are the wavelet covariant transform (WCT) PBLH (CATS WCT-PBLH), red dots are the Cloud Top Height (CTH), and magenta dots are the Cloud Base Height (CBH). (<b>b</b>) Average vertical profile of the attenuated backscatter measured by CATS (red line), and the ground-based Lidar backscatter profile measured at the same time as the CATS overpass (black line). (<b>c</b>) Distance of CATS to the SGP site in degrees. (<b>d</b>–<b>f</b>) same as (<b>a</b>–<b>c</b>) but for 29 September 2016.</p>
Full article ">Figure 5
<p>(<b>a</b>) Scatter plots showing the relationship between the ground-based Micropulse DTDS-PBLH and CATS WCT-PBLH for samples classified as good quality only (magenta dots) and samples considered good and mediate (black dots) over the Southern Great Plains (SGP). The quality classification results from the DTDS algorithm and is not incorporated in the WCT method. Classifying the WCT dataset between good (magenta dots) and good and mediate (black dots) responds to the need to make the samples of both methodologies comparable. (<b>b</b>) same as (<b>a</b>) but comparing CATS WCT-PBLH with radiosonde PBLH retrievals. (<b>c</b>,<b>d</b>) same as (<b>a</b>,<b>b</b>) but for CATS DTDS-PBLH.</p>
Full article ">Figure 6
<p>Temporal variations of the PBLH with 2 h intervals for different instruments and methodologies over the Southern Great Plains (SGP). The black line corresponds to the ground-based MPL PBLH based on DTDS (ground-based DTDS-PBLH). The magenta line is the CATS PBLH estimated using the DTDS algorithm (CATS DTDS-PBLH) only considering good-quality retrievals. The blue line is the CATS PBLH estimated using the traditional WCT method (CATS WCT-PBLH). Yellow stars are the radiosonde PBLH retrieved using the Liu and Liang method (RS-PBLH). The green line is the computed lifting condensation level (LCL) using the provided meteorological data in CATS files (CATS LCL).</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>b</b>) same as <a href="#remotesensing-16-03252-f005" class="html-fig">Figure 5</a>c,d but without the use of the lifting condensation level (LCL) in the DTDS algorithm. (<b>c</b>) same as <a href="#remotesensing-16-03252-f006" class="html-fig">Figure 6</a> but without LCL in the DTDS algorithm.</p>
Full article ">Figure 8
<p>(<b>a</b>) Scatter plots showing the relationship between the ground-based Micropulse DTDS-PBLH and CATS WCT-PBLH for samples classified as good quality only (magenta dots) and samples considered good and mediate (black dots) at the Eastern North Atlantic (ENA) site. Notice that the quality classification results from the DTDS algorithm and is not incorporated in the WCT method. Classifying the WCT dataset between good (magenta dots) and good and mediate (black dots) responds to the need to make the samples of both methodologies comparable. (<b>b</b>) same as (<b>a</b>) but comparing CATS WCT-PBLH with radiosonde PBLH retrievals. (<b>c</b>,<b>d</b>) same as (<b>a</b>,<b>b</b>) but for CATS DTDS-PBLH.</p>
Full article ">Figure 9
<p>Temporal variations of the PBLH with 2 h intervals for different instruments and methodologies over the Eastern North Atlantic (ENA). The black line corresponds to the ground-based MPL PBLH based on DTDS. The magenta line is the CATS PBLH estimated using the DTDS algorithm (CATS DTDS-PBLH). The blue line is the CATS PBLH estimated using the WCT method (also called the traditional method here). Yellow stars are the radiosonde PBLH retrieved using the Liu and Liang method.</p>
Full article ">
24 pages, 4877 KiB  
Article
A Comparative Analysis of Remote Sensing Estimation of Aboveground Biomass in Boreal Forests Using Machine Learning Modeling and Environmental Data
by Jie Song, Xuelu Liu, Samuel Adingo, Yanlong Guo and Quanxi Li
Sustainability 2024, 16(16), 7232; https://doi.org/10.3390/su16167232 - 22 Aug 2024
Viewed by 789
Abstract
It is crucial to have precise and current maps of aboveground biomass (AGB) in boreal forests to accurately track global carbon levels and develop effective plans for addressing climate change. Remote sensing as a cost-effective tool offers the potential to update AGB maps [...] Read more.
It is crucial to have precise and current maps of aboveground biomass (AGB) in boreal forests to accurately track global carbon levels and develop effective plans for addressing climate change. Remote sensing as a cost-effective tool offers the potential to update AGB maps for boreal forests in real time. This study evaluates different machine learning algorithms, namely Light Gradient Boosting Machine (LightGBM), Extreme Gradient Boosting (XGBoost), Random Forest (RF), and Support Vector Regression (SVR), for predicting AGB in boreal forests. Conducted in the Qilian Mountains, northwest China, the study integrated field measurements, space-borne LiDAR, optical remote sensing, and environmental data to develop a training dataset. Among 34 variables, 22 were selected for AGB estimation modeling. Our findings revealed that the LightGBM AGB model had the highest level of accuracy (R2 = 0.84, RMSE = 15.32 Mg/ha), outperforming the XGBoost, RF, and SVR AGB models. Notably, the LightGBM AGB model effectively addressed issues of underestimation and overestimation. We also observed that the disparity in accuracy among the models widens with increasing altitude. Remarkably, the LightGBM AGB model consistently demonstrates optimal performance across all elevation gradients, with residuals generally below 25 Mg/ha for low-value overestimation and below −38 Mg/ha for high-value underestimation. The model developed in this study presents a viable and alternative approach for enhancing AGB estimation accuracy in boreal forests based on remote sensing technology. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The arrangement of field survey plots and GLAS footprints across the study area. Representative instances of surveyed forests are illustrated in (<b>b</b>) for a Qilian mountain forest landscape and in (<b>c</b>) for a <span class="html-italic">Picea crassifolia</span> forest.</p>
Full article ">Figure 2
<p>Land cover map for the study area.</p>
Full article ">Figure 3
<p>The methodology we used in this study for estimating boreal forest AGB.</p>
Full article ">Figure 4
<p>Effect of topographic correction of GLAS-derived canopy heights based on different slope gradients.</p>
Full article ">Figure 5
<p>Frequency distribution of variables in the initial XGBoost run.</p>
Full article ">Figure 6
<p>Cross-validation results for each ML model (the different colors represent distinct models corresponding to the X-axis).</p>
Full article ">Figure 7
<p>Comparison of raster ratios across different value domains in predicted AGB maps and training data.</p>
Full article ">Figure 8
<p>Independent validation results for each ML model.</p>
Full article ">Figure 8 Cont.
<p>Independent validation results for each ML model.</p>
Full article ">Figure 9
<p>Distribution of residuals of AGB maps estimated by each model for (<b>a</b>) different value ranges and (<b>b</b>) different elevations (the red line represents no difference between the model’s predicted values and the observed values).</p>
Full article ">Figure 10
<p>Distribution of forest AGB across the study area.</p>
Full article ">
23 pages, 5725 KiB  
Article
Estimation of the Aboveground Carbon Storage of Dendrocalamus giganteus Based on Spaceborne Lidar Co-Kriging
by Huanfen Yang, Zhen Qin, Qingtai Shu, Lei Xi, Cuifen Xia, Zaikun Wu, Mingxing Wang and Dandan Duan
Forests 2024, 15(8), 1440; https://doi.org/10.3390/f15081440 - 15 Aug 2024
Viewed by 1116
Abstract
Bamboo forests, as some of the integral components of forest ecosystems, have emerged as focal points in forestry research due to their rapid growth and substantial carbon sequestration capacities. In this paper, satellite-borne lidar data from GEDI and ICESat-2/ATLAS are utilized as the [...] Read more.
Bamboo forests, as some of the integral components of forest ecosystems, have emerged as focal points in forestry research due to their rapid growth and substantial carbon sequestration capacities. In this paper, satellite-borne lidar data from GEDI and ICESat-2/ATLAS are utilized as the main information sources, with Landsat 9 and DEM data as covariates, combined with 51 pieces of ground-measured data. Using random forest regression (RFR), boosted regression tree (BRT), k-nearest neighbor (KNN), Cubist, extreme gradient boosting (XGBoost), and Stacking-ridge regression (RR) machine learning methods, an aboveground carbon (AGC) storage model was constructed at a regional scale. The model evaluation indices were the coefficient of determination (R2), root mean square error (RMSE), and overall estimation accuracy (P). The results showed that (1) The best-fit semivariogram models for cdem, fdem, fndvi, pdem, and andvi were Gaussian models, while those for h1b7, h2b7, h3b7, and h4b7 were spherical models; (2) According to Pearson correlation analysis, the AGC of Dendrocalamus giganteus showed an extremely significant correlation (p < 0.01) with cdem and pdem from GEDI, and also showed an extremely significant correlation with andvi, h1b7, h2b7, h3b7, and h4b7 from ICESat-2/ATLAS; moreover, AGC showed a significant correlation (0.01 < p < 0.05) with fdem and fndvi from GEDI; (3) The estimation accuracy of the GEDI model was superior to that of the ICESat-2/ATLAS model; additionally, the estimation accuracy of the Stacking-RR model, which integrates GEDI and ICESat-2/ATLAS (R2 = 0.92, RMSE = 5.73 Mg/ha, p = 86.19%), was better than that of any single model (XGBoost, RFR, BRT, KNN, Cubist); (4) Based on the Stacking-RR model, the estimated AGC of Dendrocalamus giganteus within the study area was 1.02 × 107 Mg. The average AGC was 43.61 Mg/ha, with a maximum value of 76.43 Mg/ha and a minimum value of 15.52 Mg/ha. This achievement can serve as a reference for estimating other bamboo species using GEDI and ICESat-2/ATLAS remote sensing technologies and provide decision support for the scientific operation and management of Dendrocalamus giganteus. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area: (<b>a</b>) The study area is located in southwest China, (<b>b</b>) Xinping is part of Yunnan Province, and (<b>c</b>) Xinping DEM, the red circle, is a collection of 51 <span class="html-italic">Dendrocalamus giganteus</span> plots.</p>
Full article ">Figure 2
<p>Schematic diagram of GEDI and ICESat-2/ATLAS spots in the study area: (<b>a</b>) GEDI spots, (<b>b</b>) ICESat-2/ATLAS spots.</p>
Full article ">Figure 3
<p>Landsat 9: (<b>a</b>) <span class="html-italic">b7</span>, (<b>b</b>) <span class="html-italic">ndvi</span>. Note: b7: Short-Wave Infrared 2. ndvi: Normalized vegetation index.</p>
Full article ">Figure 4
<p>Technology roadmap.</p>
Full article ">Figure 5
<p>Model technology roadmap.</p>
Full article ">Figure 6
<p>Correlation matrix between AGC and remote sensing factors.</p>
Full article ">Figure 7
<p>Interpolation result: (<b>a</b>) is <span class="html-italic">c<sub>dem</sub></span>, (<b>b</b>) <span class="html-italic">f<sub>dem</sub></span>, (<b>c</b>) <span class="html-italic">f<sub>ndvi</sub></span>, (<b>d</b>) <span class="html-italic">p<sub>dem</sub></span>, (<b>e</b>) <span class="html-italic">a<sub>ndvi</sub></span>, (<b>f</b>) <span class="html-italic">h1<sub>b7</sub></span>, (<b>g</b>) <span class="html-italic">h2<sub>b7</sub></span>, (<b>h</b>) <span class="html-italic">h3<sub>b7</sub></span>, and (<b>i</b>) <span class="html-italic">h4<sub>b7</sub></span>.</p>
Full article ">Figure 8
<p>Scatter diagram of the AGC model of <span class="html-italic">Dendrocalamus giganteus.</span> The models are RFR, BRT, KNN, Cubist, XGBoost, and Stacking-RR. GEDI model (<b>a</b>–<b>f</b>), ICESat-2/ATLAS model (<b>g</b>–<b>l</b>), integrated GEDI and ICESat-2/ATLAS model (<b>m</b>–<b>r</b>).</p>
Full article ">Figure 9
<p>Distribution map of <span class="html-italic">Dendrocalamus giganteus</span> AGC in Xinping County.</p>
Full article ">
22 pages, 9422 KiB  
Article
Seasonal Variability in the Relationship between the Volume-Scattering Function at 180° and the Backscattering Coefficient Observed from Spaceborne Lidar and Biogeochemical Argo (BGC-Argo) Floats
by Miao Sun, Peng Chen, Zhenhua Zhang and Yunzhou Li
Remote Sens. 2024, 16(15), 2704; https://doi.org/10.3390/rs16152704 - 24 Jul 2024
Viewed by 662
Abstract
The derivation of the particulate-backscattering coefficient (bbp) from Lidar signals is highly influenced by the parameter χp(π), which is defined by χp(π) = bbp/(2πβp(π)). This parameter facilitates the correlation of the [...] Read more.
The derivation of the particulate-backscattering coefficient (bbp) from Lidar signals is highly influenced by the parameter χp(π), which is defined by χp(π) = bbp/(2πβp(π)). This parameter facilitates the correlation of the particulate-volume-scattering function at 180°, denoted βp(π), with bbp. However, studies exploring the global and seasonal fluctuations of χp(π) remain sparse, largely due to measurement difficulties of βp(π) in the field conditions. This study pioneers the global data collection for χp(π), integrating bbp observations from Biogeochemical Argo (BGC-Argo) floats and βp(π) data from the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP) spaceborne lidar. Our findings indicate that χp(π) experiences significant seasonal differences globally, peaking during summer and nadiring in winter. The global average χp(π) was calculated as 0.40, 0.48, 0.43, and 0.35 during spring, summer, autumn, and winter, respectively. The daytime values of χp(π) slightly exceeded those registered at night. To illuminate the seasonal variations in χp(π) in 26 sea regions worldwide, we deployed passive ocean color data MODIS bbp and active remote sensing data CALIOP βp(π), distinguishing three primary seasonal change patterns—the “summer peak”, the “decline”, and the “autumn pole”—with the “summer peak” typology being the most common. Post recalibration of the CALIOP bbp product considering seasonal χp(π) variations, we observed substantial statistical improvements. Specifically, the coefficient of determination (R2) markedly improved from 0.84 to 0.89, while the root mean square error (RMSE) declined from 4.0 × 10−4 m−1 to 3.0 × 10−4 m−1. Concurrently, the mean absolute percentage error (MAPE) also dropped significantly, from 31.48% to 25.27%. Full article
(This article belongs to the Section Environmental Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>A comparison of spatial distribution and quantity of BGC-Argo floats after quality control. (<b>a</b>) illustrates the spatial distribution of BGC-Argo buoys; (<b>b</b>) shows the comparison of the number of buoys with different quality control factors.</p>
Full article ">Figure 2
<p>The results for matching points of BGC-Argo and CALIOP using different spatial windows, including a 9 km window (<b>a</b>), a 50 km window (<b>b</b>), and a 1° × 1° window (<b>c</b>). (<b>d</b>) shows a comparison of the number of matching points across different spatiotemporal windows.</p>
Full article ">Figure 3
<p>The comparison between CALIOP and BGC-Argo <span class="html-italic">b</span><sub>bp</sub>(532 nm) matchups for a 9 km spatial window and a ±3 h time window, as well as the comparison between CALIOP and MODIS <span class="html-italic">b</span><sub>bp</sub> (532 nm) matchups for a 9 km spatial window and a ±12 h time window. (<b>a</b>,<b>b</b>) denote the fitting results of CALIOP <span class="html-italic">b</span><sub>bp</sub>(532) using the <span class="html-italic">χ</span><sub>p</sub>(π) = 1.00 algorithm and BGC-Argo <span class="html-italic">b</span><sub>bp</sub>(532) estimates matching; (<b>c</b>,<b>d</b>) denote the fitting results of CALIOP <span class="html-italic">b</span><sub>bp</sub>(532) using the <span class="html-italic">χ</span><sub>p</sub>(π) = 0.50 algorithm and BGC-Argo <span class="html-italic">b</span><sub>bp</sub>(532) estimates matching; and (<b>e</b>,<b>f</b>) denote the fitting results of CALIOP <span class="html-italic">b</span><sub>bp</sub>(532) using the <span class="html-italic">χ</span><sub>p</sub>(π) = 0.50 algorithm and MODIS <span class="html-italic">b</span><sub>bp</sub>(532) estimates matching.</p>
Full article ">Figure 4
<p>The flow chart of this study.</p>
Full article ">Figure 5
<p>The comparison of <span class="html-italic">χ</span><sub>p</sub>(π) values of conversion coefficients for different seasons within different spatial–temporal matching windows. (<b>a</b>) shows the <span class="html-italic">χ</span><sub>p</sub>(π) calibration results with a 9 km spatial window match; (<b>b</b>) shows the calibration results with a 50 km spatial window match; and (<b>c</b>) shows the calibration results with a 1° × 1° spatial window match.</p>
Full article ">Figure 6
<p>Line plots of the seasonal variation of <span class="html-italic">χ</span><sub>p</sub>(π) for 12 spatiotemporally matched windows. (<b>a</b>) A line graph representing the seasonal variation of <span class="html-italic">χ</span><sub>p</sub>(π) for each spatial–temporal window; (<b>b</b>) A bar graph representing the seasonal variation of the mean <span class="html-italic">χ</span><sub>p</sub>(π).</p>
Full article ">Figure 7
<p>Variations of <span class="html-italic">χ</span><sub>p</sub>(π) values between day and night for different seasons under a 1° × 1° spatial matching window. (<b>a</b>) presents comparisons of <span class="html-italic">χ</span><sub>p</sub>(π) calibration results between day and night for various seasons with a ±12 h time matching window; (<b>b</b>) shows comparisons of <span class="html-italic">χ</span><sub>p</sub>(π) calibration results between day and night for various seasons with a ±24 h time matching window; (<b>c</b>) displays bar graphs illustrating comparisons of <span class="html-italic">χ</span><sub>p</sub>(π) calibration results between day and night for various seasons with a ±12 h time matching window; and (<b>d</b>) presents bar graphs for comparisons of <span class="html-italic">χ</span><sub>p</sub>(π) calibration results between day and night for various seasons with a ±24 h time matching window.</p>
Full article ">Figure 8
<p>Performance comparison of CALIOP <span class="html-italic">b</span><sub>bp</sub> products before and after calibration in a 9 km spatial window. (<b>a</b>) R<sup>2</sup>, (<b>b</b>) RMSE, (<b>c</b>) MAPE, and (<b>d</b>) SD Based on BGC-Argo Evaluation.</p>
Full article ">Figure 9
<p>MODIS-corrected <span class="html-italic">χ</span><sub>p</sub>(π) seasonal line plots for various sea areas around the globe.</p>
Full article ">
22 pages, 9521 KiB  
Article
Estimation of Leaf Area Index for Dendrocalamus giganteus Based on Multi-Source Remote Sensing Data
by Zhen Qin, Huanfen Yang, Qingtai Shu, Jinge Yu, Li Xu, Mingxing Wang, Cuifen Xia and Dandan Duan
Forests 2024, 15(7), 1257; https://doi.org/10.3390/f15071257 - 19 Jul 2024
Viewed by 1321
Abstract
The Leaf Area Index (LAI) plays a crucial role in assessing the health of forest ecosystems. This study utilized ICESat-2/ATLAS as the primary information source, integrating 51 measured sample datasets, and employed the Sequential Gaussian Conditional Simulation (SGCS) method to derive surface grid [...] Read more.
The Leaf Area Index (LAI) plays a crucial role in assessing the health of forest ecosystems. This study utilized ICESat-2/ATLAS as the primary information source, integrating 51 measured sample datasets, and employed the Sequential Gaussian Conditional Simulation (SGCS) method to derive surface grid information for the study area. The backscattering coefficient and texture feature factor from Sentinel-1, as well as the spectral band and vegetation index factors from Sentinel-2, were integrated. The random forest (RF), gradient-boosted regression tree (GBRT) model, and K-nearest neighbor (KNN) method were employed to construct the LAI estimation model. The optimal model, RF, was selected to conduct accuracy analysis of various remote sensing data combinations. The spatial distribution map of Dendrocalamus giganteus in Xinping County was then generated using the optimal combination model. The findings reveal the following: (1) Four key parameters—optimal fitted segmented terrain height, interpolated terrain surface height, absolute mean canopy height, and solar elevation angle—are significantly correlated. (2) The RF model constructed using a combination of ICESat-2/ATLAS, Sentinel-1, and Sentinel-2 data achieved optimal accuracy, with a coefficient of determination (R2) of 0.904, root mean square error (RMSE) of 0.384, mean absolute error (MAE) of 0.319, overall estimation accuracy (P1) of 88.96%, and relative root mean square error (RRMSE) of 11.04%. (3) The accuracy of LAI estimation using a combination of ICESat-2/ATLAS, Sentinel-1, and Sentinel-2 remote sensing data showed slight improvement compared to using either ICESat-2/ATLAS data combined with Sentinel-1 or Sentinel-2 data alone, with a significant enhancement in LAI estimation accuracy compared to using ICESat-2/ATLAS data alone. (4) LAI values in the study area ranged mainly from 2.29 to 2.51, averaging 2.4. Research indicates that employing ICESat-2/ATLAS spaceborne LiDAR data for regional-scale LAI estimation presents clear advantages. Incorporating SAR data and optical imagery and utilizing diverse data types for complementary information significantly enhances the accuracy of LAI estimation, demonstrating the feasibility of LAI inversion with multi-source remote sensing data. This approach offers an innovative framework for utilizing multi-source remote sensing data for regional-scale LAI inversion, demonstrates a methodology for integrating various remote sensing data, and serves as a reference for low-cost high-precision regional-scale LAI estimation. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Xinping Yi and Dai Autonomous County. Note: The base map in this figure is named “Administrative Map of Yunnan Province”, with the review number Yun S (2020) No. 102. It was supervised by the Yunnan Provincial Department of Natural Resources and produced by the Yunnan Provincial Map Institute. The same applies below.</p>
Full article ">Figure 2
<p>Technology roadmap.</p>
Full article ">Figure 3
<p>Light spot distribution map in the study area.</p>
Full article ">Figure 4
<p>SGCS effect: (<b>a</b>) h_te_best_fit; (<b>b</b>) h_te_interp; (<b>c</b>) h_mean_canopy_abs; and (<b>d</b>) solar_elevation.</p>
Full article ">Figure 5
<p>Correlation coefficient thermal matrix diagram.</p>
Full article ">Figure 6
<p>Model fitting scatter plot: (<b>a</b>) RF; (<b>b</b>) GBRT; and (<b>c</b>) KNN.</p>
Full article ">Figure 7
<p>Using only ICESat-2/ATLAS data: (<b>a</b>) RF; (<b>b</b>) GBRT; and (<b>c</b>) KNN.</p>
Full article ">Figure 8
<p>ICESat-2/ATLAS and Sentinel-1 combination: (<b>a</b>) RF; (<b>b</b>) GBRT; and (<b>c</b>) KNN.</p>
Full article ">Figure 9
<p>ICESat-2/ATLAS and Sentinel-2 combination: (<b>a</b>) RF; (<b>b</b>) GBRT; and (<b>c</b>) KNN.</p>
Full article ">Figure 10
<p>Spatial distribution map of <span class="html-italic">Dendrocalamus giganteus</span> LAI in the study area.</p>
Full article ">
28 pages, 14496 KiB  
Article
An Optimal Denoising Method for Spaceborne Photon-Counting LiDAR Based on a Multiscale Quadtree
by Baichuan Zhang, Yanxiong Liu, Zhipeng Dong, Jie Li, Yilan Chen, Qiuhua Tang, Guoan Huang and Junlin Tao
Remote Sens. 2024, 16(13), 2475; https://doi.org/10.3390/rs16132475 - 5 Jul 2024
Cited by 1 | Viewed by 1095
Abstract
Ice, Cloud, and Land Elevation Satellite-2 (ICESat-2) has excellent potential for obtaining water depth information around islands and reefs. Combining the density-based spatial clustering of applications with noise algorithm (DBSCAN) and multiscale quadtree analysis, we propose a new photon-counting lidar denoising method to [...] Read more.
Ice, Cloud, and Land Elevation Satellite-2 (ICESat-2) has excellent potential for obtaining water depth information around islands and reefs. Combining the density-based spatial clustering of applications with noise algorithm (DBSCAN) and multiscale quadtree analysis, we propose a new photon-counting lidar denoising method to discard the large amount of noise in ICESat-2 data. First, the kernel density estimation (KDE) is used to preprocess the point cloud data, and a threshold is set to remove the noise photons on the sea surface. Next, the DBSCAN algorithm is used to preliminarily remove underwater noise photons. Then, the quadtree segmentation and Otsu algorithm are used for fine denoising to extract accurate bottom signal photons. Based on ICESat-2 pho-ton-counting data from six typical islands and reefs worldwide, the proposed method outperforms other algorithms in terms of denoising effect. Compared to in situ data, the determination coefficient (R2) reaches 94.59%, and the root mean square error (RMSE) is 1.01 m. The proposed method can extract accurate underwater terrain information, laying a foundation for offshore bathymetry. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of Experimental Areas. (<b>a</b>) Oahu Island; (<b>b</b>) Dongdao Island; (<b>c</b>) Huaguang Reef; (<b>d</b>) Ailinginae Atoll. The solid lines in each study area represent the flight trajectory of ICESat-2, and the yellow diagonal lines in (<b>b</b>) represent the in situ depth measurement data.</p>
Full article ">Figure 2
<p>ATL03 raw data points distribution and reference data. (<b>a</b>) Distribution of the ATL03 raw data. (<b>b</b>) Example of the classification results of the reference data.</p>
Full article ">Figure 3
<p>Method flow chart.</p>
Full article ">Figure 4
<p>Extraction of sea surface photon points. (<b>a</b>) The overall KDE curve. (<b>b</b>) The photon extraction details.</p>
Full article ">Figure 5
<p>Sea surface photons. (<b>a</b>) The distribution of the sea surface photons. (<b>b</b>) The extraction of the sea surface signal photons.</p>
Full article ">Figure 6
<p>DBSCABN clustering principle. The blue circles represent the signal photons, and the red circle represents the noise photons.</p>
Full article ">Figure 7
<p>Quadtree flowchart. (<b>a</b>) The quadtree segmentation process, with four segmentation layers and a maximum spatial photon capacity of 20. (<b>b</b>) The quadtree structure.</p>
Full article ">Figure 8
<p>Optimized quadtree flowchart. (<b>a</b>) The optimized quadtree segmentation process, with four segmentation layers and a maximum spatial photon capacity of 20. (<b>b</b>) The optimized quadtree structure.</p>
Full article ">Figure 9
<p>Experimental results. (<b>a1</b>–<b>a3</b>) Oahu Island; (<b>b1</b>–<b>b3</b>) Dongdao Island in 10 September 2022; (<b>c1</b>–<b>c3</b>) Dongdao Island in 7 August 2023; (<b>d1</b>–<b>d3</b>) Huaguang Reef in 14 January 2022; (<b>e1</b>–<b>e3</b>) Huaguang Reef in 15 July 2022; (<b>f1</b>–<b>f3</b>) Ailinginae Atoll Island.</p>
Full article ">Figure 9 Cont.
<p>Experimental results. (<b>a1</b>–<b>a3</b>) Oahu Island; (<b>b1</b>–<b>b3</b>) Dongdao Island in 10 September 2022; (<b>c1</b>–<b>c3</b>) Dongdao Island in 7 August 2023; (<b>d1</b>–<b>d3</b>) Huaguang Reef in 14 January 2022; (<b>e1</b>–<b>e3</b>) Huaguang Reef in 15 July 2022; (<b>f1</b>–<b>f3</b>) Ailinginae Atoll Island.</p>
Full article ">Figure 10
<p>Comparison of Dongdao Island denoising results in 10 September 2022. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 11
<p>Comparison of Huaguang Reef denoising results in 15 July 2022. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 12
<p>Comparison of Oahu Island denoising results. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 12 Cont.
<p>Comparison of Oahu Island denoising results. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 13
<p>Comparison of Dongdao Island denoising results in 7 August 2023. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 13 Cont.
<p>Comparison of Dongdao Island denoising results in 7 August 2023. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 14
<p>Comparison of Huaguang Reef denoising results in 14 January 2022. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 15
<p>Comparison of Huaguang Reef denoising results. (<b>a</b>) ATL03 results. (<b>b</b>) DBSCAN results. (<b>c</b>) Quadtree results. (<b>d</b>) Results of the proposed method.</p>
Full article ">Figure 16
<p>Scatterplots of ICESat-2 extracted depth and in situ depth for different algorithms. (<b>a</b>) DBSCAN; (<b>b</b>) Quadtree; (<b>c</b>) The proposed method.</p>
Full article ">Figure 17
<p>Comparison between the ICESat-2 denoising results and in situ data.</p>
Full article ">Figure 18
<p>Quantitative evaluation of the performances of the four methods at different water depths. (<b>a</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m in Oahu Island. (<b>b</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m on 10 September 2022 in Dongdao Island. (<b>c</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m on 7 August 2023 in Dongdao Island. (<b>d</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m on 14 January 2022 in Huaguang Reef. (<b>e</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m on 15 July 2022 in Huaguang Reef. (<b>f</b>) Represent the result of &lt;10 m, 10–20 m, and &gt;20 m in Ailinginae Atoll.</p>
Full article ">
Back to TopTop