[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
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (163)

Search Parameters:
Keywords = linear unmixing

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
9 pages, 3733 KiB  
Communication
Alkamide Content and Localization in Heliopsis longipes Cypselae, Obtained via Fluorescence and Double-Multiphoton Microscopy
by Juan Vázquez-Martínez and Jorge Molina-Torres
Molecules 2024, 29(23), 5651; https://doi.org/10.3390/molecules29235651 - 29 Nov 2024
Viewed by 521
Abstract
The alkamide content and specific tissue localization in the cypselae of Heliopsis longipes were investigated using gas chromatography–electron ionization mass spectrometry (GC-EIMS) and multiphoton fluorescence microscopy (MPFM). GC-EIMS analysis identified two olefinic alkamides: affinin (spilanthol) and N-2-methylbutyl-2E,6Z,8E [...] Read more.
The alkamide content and specific tissue localization in the cypselae of Heliopsis longipes were investigated using gas chromatography–electron ionization mass spectrometry (GC-EIMS) and multiphoton fluorescence microscopy (MPFM). GC-EIMS analysis identified two olefinic alkamides: affinin (spilanthol) and N-2-methylbutyl-2E,6Z,8E-decatrienamide. Microscopic analysis revealed that alkamides are localized within the cotyledons, and specifically compartmentalized in lipid bodies, highlighting their spatial organization. The linear unmixing of fluorescence emission fingerprints showed that affinin exhibits autofluorescence at 693 nm, corresponding to the red spectral region. This emission is attributed to the conjugated double bonds in its acyl chain. This study is the first to report on the presence and precise localization of alkamides in the cypselae of H. longipes. Full article
Show Figures

Figure 1

Figure 1
<p>Structures of <span class="html-italic">H. longipes</span> cypselae (Cy) used for alkamides analyses: pericarp (Pr), seedcoat more cotyledon (Sc+), seedcoat (Sc), and cotyledon (Cot).</p>
Full article ">Figure 2
<p>GC-EIMS analysis of <span class="html-italic">H. longipes</span> Cot ethanolic extract. (<b>A</b>) Total ion chromatogram of Cot extract; the components retained at *18.785, *21.280, *23.334 min are affinin, <span class="html-italic">N</span>-2-methylbutyl-2<span class="html-italic">E</span>,6<span class="html-italic">Z</span>,8<span class="html-italic">E</span>-decatrienamide, and farnesol acetate, respectively. (<b>B</b>) <span class="html-italic">N</span>-isobutyl-<span class="html-italic">2E,6Z,8E</span>-decatrienamide (affinin) mass spectrum; comparison between the component retained at 18.785 (red) and the authentic compound (blue). (<b>C</b>) <span class="html-italic">N</span>-2-methylbutyl-2<span class="html-italic">E</span>,6<span class="html-italic">Z</span>,8<span class="html-italic">E</span>-decatrienamide mass spectrum; comparison between the component retained at 21.280 (red) and the authentic compound (blue).</p>
Full article ">Figure 3
<p>Transversal slice of <span class="html-italic">H. longipes</span> cypsela stained with NR (NR) observed using an optical microscope (10×). Cot = cotyledon; Ms = seedcoat; Ex = pericarp.</p>
Full article ">Figure 4
<p>Lambda stack emission spectra: (<b>A</b>) affinin; (<b>B</b>) affinin + Nile Red; (<b>C</b>) <span class="html-italic">N</span>-isobutyl-decanamide.</p>
Full article ">Figure 5
<p>Autofluorescence of selected compounds present in a transversal slice of <span class="html-italic">H. longipes</span> cotyledons observed under multiphoton microscope. On the left, triglycerides stained with Nile Red. In the middle, the autofluorescence of alkamides (affinin). On the right, the overlapping of the two previous results.</p>
Full article ">
19 pages, 7749 KiB  
Article
Generative Simplex Mapping: Non-Linear Endmember Extraction and Spectral Unmixing for Hyperspectral Imagery
by John Waczak and David J. Lary
Remote Sens. 2024, 16(22), 4316; https://doi.org/10.3390/rs16224316 - 19 Nov 2024
Viewed by 704
Abstract
We introduce a new model for non-linear endmember extraction and spectral unmixing of hyperspectral imagery called Generative Simplex Mapping (GSM). The model represents endmember mixing using a latent space of points sampled within a (n1)-simplex corresponding to n [...] Read more.
We introduce a new model for non-linear endmember extraction and spectral unmixing of hyperspectral imagery called Generative Simplex Mapping (GSM). The model represents endmember mixing using a latent space of points sampled within a (n1)-simplex corresponding to n unique sources. Barycentric coordinates within this simplex are naturally interpreted as relative endmember abundances satisfying both the abundance sum-to-one and abundance non-negativity constraints. Points in this latent space are mapped to reflectance spectra via a flexible function combining linear and non-linear mixing. Due to the probabilistic formulation of the GSM, spectral variability is also estimated by a precision parameter describing the distribution of observed spectra. Model parameters are determined using a generalized expectation-maximization algorithm, which guarantees non-negativity for extracted endmembers. We first compare the GSM against three varieties of non-negative matrix factorization (NMF) on a synthetic data set of linearly mixed spectra from the USGS spectral database. Here, the GSM performed favorably for both endmember accuracy and abundance estimation with all non-linear contributions driven to zero by the fitting procedure. In a second experiment, we apply the GTM to model non-linear mixing in real hyperspectral imagery captured over a pond in North Texas. The model accurately identified spectral signatures corresponding to near-shore algae, water, and rhodamine tracer dye introduced into the pond to simulate water contamination by a localized source. Abundance maps generated using the GSM accurately track the evolution of the dye plume as it mixes into the surrounding water. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of the non-linear activation functions for a 2-component GSM. The simplex vertices are shown as green dots (<math display="inline"><semantics> <msub> <mi>μ</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>μ</mi> <mn>2</mn> </msub> </semantics></math>) and three RBF centers are shown in red (<math display="inline"><semantics> <msub> <mi>μ</mi> <mn>1</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>μ</mi> <mn>2</mn> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mi>μ</mi> <mn>3</mn> </msub> </semantics></math>). With no RBF centers at the vertices, we guarantee no non-linear contributions occur for pure endmember spectra.</p>
Full article ">Figure 2
<p>Illustration of the GSM. The latent space consists of a grid of <span class="html-italic">K</span>-many points (green dots) distributed throughout a simplex with <math display="inline"><semantics> <msub> <mi>N</mi> <mi>v</mi> </msub> </semantics></math> vertices. Barycentric coordinates of each node in the simplex correspond to the relative abundance of <math display="inline"><semantics> <msub> <mi>N</mi> <mi>v</mi> </msub> </semantics></math>-many unique sources. Here, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mi>v</mi> </msub> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> has been chosen for illustrative purposes. Nodes are mapped into the data space via the map <math display="inline"><semantics> <mrow> <mi>ψ</mi> <mo>(</mo> <mi mathvariant="bold">z</mi> <mo>)</mo> </mrow> </semantics></math> utilizing <span class="html-italic">M</span>-many radially symmetric basis functions (red). Spectral variability is estimated via the precision parameter, <math display="inline"><semantics> <mi>β</mi> </semantics></math>, shown here in the data space as a light blue band around the spectrum given by <math display="inline"><semantics> <mrow> <mi>ψ</mi> <mo>(</mo> <mi mathvariant="bold">z</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Synthetic data set formed from USGS spectra. (<b>a</b>) Spectra from the USGS spectral database used as the ground truth endmembers. These spectra were selected following the example in ref. [<a href="#B13-remotesensing-16-04316" class="html-bibr">13</a>]. (<b>b</b>) The abundance distribution sampled for in the data set. Samples were generated from a Dirichlet distribution with <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>1</mn> </msub> <mo>=</mo> <msub> <mi>α</mi> <mn>2</mn> </msub> <mo>=</mo> <msub> <mi>α</mi> <mn>3</mn> </msub> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Real HSI data set. (<b>a</b>) The UAV used to collect hyperspectral images. (<b>b</b>) The Resonon Pika XC2 hyperspectral imager used to acquire HSI. (<b>c</b>) A sample hyperspectral data cube. Spectra are plotted for each pixel at their geographic position. The log10-reflectance is colored along the z axis, and a pseudocolor image is shown on top of the data cube. The signature of the rhodamine dye plume is clearly identifiable in the water.</p>
Full article ">Figure 5
<p>Explained variance of PCA components for the real HSI data set. A red horizontal line is superimposed on the graph, marking an explained variance of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math>. All components past the fourth explain less than <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math> of the observed variance.</p>
Full article ">Figure 6
<p>Comparison of GSM against NMF on simulated linear mixing data set using USGS spectra. (<b>a</b>) The mean spectral angle computed between extracted endmembers and original endmembers. (<b>b</b>) The mean RMSE computed between extracted endmembers and original endmembers. (<b>c</b>) The mean abundance RMSE computed between original abundance data for each endmember and extracted abundances. (<b>d</b>) The reconstruction RMSE, which evaluates the quality of fit. All models realized similar values, reflecting convergence of the models to the level of random noise introduced into the data.</p>
Full article ">Figure 7
<p>Endmembers extracted by the GSM for the simulated linear mixing data set with SNR <math display="inline"><semantics> <mrow> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>. The dashed lines correspond to original endmember spectra from the USGS spectral database. Solid lines superimposed on the plot indicate the extracted endmember spectra. Colored bands are included around each spectrum corresponding to the spectral variability estimated by the GSM precision parameter <math display="inline"><semantics> <mi>β</mi> </semantics></math> where the band width is <math display="inline"><semantics> <mrow> <mn>2</mn> <msqrt> <msup> <mi>β</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </msqrt> </mrow> </semantics></math> corresponding to 2 standard deviations.</p>
Full article ">Figure 8
<p>GSM applied to water spectra from real HSI data set: (<b>a</b>) Spectra generated by the trained GSM for samples with maximum abundance for each endmember. Based on these spectral profiles, endmembers are identified with water, near-shore vegetation, and rhodamine dye. (<b>b</b>) An HSI segmented according to the relative abundance of each endmember. Each water pixel is colored by smoothing interpolating between red, green, and blue colors using the relative abundance estimated for rhodamine, vegetation, and water spectra. The rhodamine plume is clearly identifiable in the western portion of the HSI.</p>
Full article ">Figure 9
<p>Endmember abundance distributions: (<b>a</b>) The spatial distribution of abundance for the water class. This source dominates in the center of the pond and decreases towards the shore where vegetation begins to dominate the reflectance signal. The water endmember abundance is also observed to decrease near the edge of the rhodamine plume reflecting dye mixing and diffusion. (<b>b</b>) The spatial distribution of vegetation. This endmember includes filamentous blue-green algae observed to accumulate in shallow waters near the shore. (<b>c</b>) The rhodamine dye plume extent segmented from the HSI. The total areas for near-shore vegetation and rhodamine are estimated to be 378.6 m<sup>2</sup> and 255.7 m<sup>2</sup>, respectively.</p>
Full article ">Figure 10
<p>Rhodamine plume evolution: Using the trained GSM we can track the dispersion of the rhodamine dye plume between successive drone flights. (<b>a</b>) The initial plume distribution after release. Here, the dye subsumes an area of <math display="inline"><semantics> <mrow> <mn>255.7</mn> </mrow> </semantics></math><math display="inline"><semantics> <msup> <mi mathvariant="normal">m</mi> <mn>2</mn> </msup> </semantics></math>. (<b>b</b>) The same plume imaged 15 min later now extends across an area of <math display="inline"><semantics> <mrow> <mn>571.8</mn> </mrow> </semantics></math><math display="inline"><semantics> <msup> <mi mathvariant="normal">m</mi> <mn>2</mn> </msup> </semantics></math>.</p>
Full article ">
21 pages, 14797 KiB  
Article
A Parameter Optimized Method for InVEST Model in Sub-Pixel Scale Integrating Machine Learning Algorithm and Vegetation–Impervious Surface–Soil Model
by Linlin Wu and Fenglei Fan
Land 2024, 13(11), 1876; https://doi.org/10.3390/land13111876 - 10 Nov 2024
Viewed by 556
Abstract
The InVEST model, with its ability to perform spatial visualization and quantification, is an important tool for mapping ecosystem services. However, the spatial accuracy and simulating performance of the model are deeply influenced by the land use parameter, which often relies on the [...] Read more.
The InVEST model, with its ability to perform spatial visualization and quantification, is an important tool for mapping ecosystem services. However, the spatial accuracy and simulating performance of the model are deeply influenced by the land use parameter, which often relies on the accuracy of land use/cover data. To address this issue, we propose a novel method for optimizing the land use parameter of the InVEST model based on the vegetation–impervious surface–soil (V–I–S) model and a machine learning algorithm. The optimized model is called Sub-InVEST, and it improves the performance of assessing ecosystem services on a sub-pixel scale. The conceptual steps are (i) extracting the V–I–S fraction of remote sensing images based on the spectral unmixing method; (ii) determining the mapping relationship of the V–I–S fraction between land use/cover type using a machine learning algorithm and field observation data; (iii) inputting the V–I–S fraction into the original model instead of the land use/cover parameter of the InVEST model. To evaluate the performance and spatial accuracy of the Sub-InVEST model, we employed the habitat quality module of InVEST and multi-source remote sensing data, which were applied to acquire Sub-InVEST and estimate the habitat quality of central Guangzhou city from 2000 to 2020 with the help of the LSMA and ISODATA methods. The experimental results showed that the Sub-InVEST model is robust in assessing ecosystem services in sets of complex ground scenes. The spatial distribution of the habitat quality of both models revealed a consistent increasing trend from the southwest to the northeast. Meanwhile, linear regression analyses observed a robust correlation and consistent linear trends, with R2 values of 0.41, 0.35, 0.42, 0.39, and 0.47 for the years 2000, 2005, 2010, 2015, and 2020, respectively. Compared with the original model, Sub-InVEST had a more favorable performance in estimating habitat quality in central Guangzhou. The spatial depictions and numerical distribution of the results of the Sub-InVSET model manifest greater detail and better concordance with remote sensing imagery and show a more seamless density curve and a substantially enhanced probability distribution across interval ranges. Full article
(This article belongs to the Section Land Environmental and Policy Impact Assessment)
Show Figures

Figure 1

Figure 1
<p>The V–I–S fraction combination for a mixed pixel.</p>
Full article ">Figure 2
<p>A flowchart of optimizing the land use parameter of the InVEST model based on the V–I–S model.</p>
Full article ">Figure 3
<p>The location of the study area in (<b>a</b>) China, (<b>b</b>) Guangdong Province, and Guangzhou City, and (<b>c</b>) remote sensing imagery of the study area in 2020.</p>
Full article ">Figure 4
<p>The spatial distribution of V–I–S fractions in central Guangzhou from 2000 to 2020.</p>
Full article ">Figure 5
<p>(<b>a</b>) The spatial distribution of habitat quality based on Sub-InVEST and InVEST. (<b>b</b>) The numerical distribution of habitat quality based on Sub-InVEST and InVEST.</p>
Full article ">Figure 6
<p>The location of sample points and sample regions for comparative assessment.</p>
Full article ">Figure 7
<p>The linear fitting of the InVEST and Sub-InVEST habitat quality results in (<b>a</b>) 2000, (<b>b</b>) 2005, (<b>c</b>) 2010, (<b>d</b>) 2015, (<b>e</b>) 2020.</p>
Full article ">Figure 8
<p>The habitat quality results based on Sub-InVEST, InVEST, and Landsat imagery for (<b>a</b>) 2000 and (<b>b</b>) 2020.</p>
Full article ">Figure 9
<p>The habitat quality results based on Sub-InVEST, InVEST, and remote sensing imagery in 2020.</p>
Full article ">
25 pages, 4756 KiB  
Article
An Adaptive Unmixing Method Based on Iterative Multi-Objective Optimization for Surface Water Fraction Mapping (IMOSWFM) Using Zhuhai-1 Hyperspectral Images
by Cong Lei, Rong Liu, Zhiyuan Kuang and Ruru Deng
Remote Sens. 2024, 16(21), 4038; https://doi.org/10.3390/rs16214038 - 30 Oct 2024
Viewed by 506
Abstract
Surface water fraction mapping is an essential preprocessing step for the subpixel mapping (SPM) of surface water, providing valuable prior knowledge about surface water distribution at the subpixel level. In recent years, spectral mixture analysis (SMA) has been extensively applied to estimate surface [...] Read more.
Surface water fraction mapping is an essential preprocessing step for the subpixel mapping (SPM) of surface water, providing valuable prior knowledge about surface water distribution at the subpixel level. In recent years, spectral mixture analysis (SMA) has been extensively applied to estimate surface water fractions in multispectral images by decomposing each mixed pixel into endmembers and their corresponding fractions using linear or nonlinear spectral mixture models. However, challenges emerge when introducing existing surface water fraction mapping methods to hyperspectral images (HSIs) due to insufficient exploration of spectral information. Additionally, inaccurate extraction of endmembers can result in unsatisfactory water fraction estimations. To address these issues, this paper proposes an adaptive unmixing method based on iterative multi-objective optimization for surface water fraction mapping (IMOSWFM) using Zhuhai-1 HSIs. In IMOSWFM, a modified normalized difference water fraction index (MNDWFI) was developed to fully exploit the spectral information. Furthermore, an iterative unmixing framework was adopted to dynamically extract high-quality endmembers and estimate their corresponding water fractions. Experimental results on the Zhuhai-1 HSIs from three test sites around Nanyi Lake indicate that water fraction maps obtained by IMOSWFM are closest to the reference maps compared with the other three SMA-based surface water fraction estimation methods, with the highest overall accuracy (OA) of 91.74%, 93.12%, and 89.73% in terms of pure water extraction and the lowest root-mean-square errors (RMSE) of 0.2506, 0.2403, and 0.2265 in terms of water fraction estimation. This research provides a reference for adapting existing surface water fraction mapping methods to HSIs. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of study areas and the corresponding Zhuhai-1 OHS false color images consisting of R-band 21, G-band 10, and B-band 7.</p>
Full article ">Figure 2
<p>Framework of IMOSWFM.</p>
Full article ">Figure 3
<p>Distribution of the selected solution in the normalized objective space.</p>
Full article ">Figure 4
<p>Spectra of typical ground features in Zhuhai-1 OHS image.</p>
Full article ">Figure 5
<p>Illustration of the double threshold method from the histogram of MNDWFI.</p>
Full article ">Figure 6
<p>Sketch map of the iterative estimation of water fraction in IMOSWFM.</p>
Full article ">Figure 7
<p>Surface water fraction maps of Area 1, Area 2, and Area 3 from IMOSWFM and other compared methods with Zhuhai-1 OHS images.</p>
Full article ">Figure 8
<p>Surface water fraction maps of Area 4, Area 5, and Area 6 from IMOSWFM and other compared methods with Zhuhai-1 OHS images.</p>
Full article ">Figure 9
<p>Comparison of NDWI and NDWFI map of Area 4. (<b>a</b>) Zhuhai-1 OHS false color image of Area 4; (<b>b</b>) NDWI map of Area 4; (<b>c</b>) MNDWFI map of Area 4; (<b>d</b>) Spectra of representative pixels in Area 4.</p>
Full article ">Figure 10
<p>Histograms of NDWI, NDWFI, and MNDWFI on the three areas around Nanyi Lake.</p>
Full article ">Figure 11
<p>RMSE and SE of the four different combinations of components on the three areas. (<b>a</b>) RMSE; (<b>b</b>) SE.</p>
Full article ">Figure 12
<p>Convergence curves of objective function values at different iterations. First row: convergence curves of the volume inverse; Second row: convergence curves of RMSE; (<b>a</b>–<b>e</b>) represent the corresponding iteration number from 1 to 5.</p>
Full article ">Figure 13
<p>Variation in the number of remaining mixed pixels as iteration increases.</p>
Full article ">
20 pages, 12465 KiB  
Article
Status, Sources, and Risks of Heavy Metals in Surface Sediments of Baiyangdian Lake and Inflow Rivers, North China
by Hongwei Liu, Yaonan Bai, Yihang Gao, Bo Han, Jinjie Miao, Yanchao Shi and Fengtian Yang
Water 2024, 16(19), 2723; https://doi.org/10.3390/w16192723 - 25 Sep 2024
Viewed by 938
Abstract
Baiyangdian Lake, recognized as the largest freshwater body in northern China, plays a vital role in maintaining the regional eco-environment. Prior studies have pointed out the contamination of sediments with heavy metals, raising concerns about eco-environmental challenges. Therefore, it is imperative to evaluate [...] Read more.
Baiyangdian Lake, recognized as the largest freshwater body in northern China, plays a vital role in maintaining the regional eco-environment. Prior studies have pointed out the contamination of sediments with heavy metals, raising concerns about eco-environmental challenges. Therefore, it is imperative to evaluate the current pollution levels and ecological threats related to heavy metals found in the sediments of Baiyangdian Lake as well as in its inflow rivers. In May 2022, surface sediments with a depth of less than 20 cm were analyzed for Cu, Zn, Pb, Cr, Ni, As, Cd, and Hg to determine the pollution status, identify sources of pollution, and evaluate potential ecological risks. A range of evaluation methods used by predecessors such as geo-accumulation index (Igeo), enrichment factor (EF), ecological risk index (RI), sediment quality guidelines (SQGs), positive matrix factorization (PMF), absolute principal component score-multiple linear regression model (APCS-MLR), chemical mass balance (CMB), and UNMIX model were analyzed. After comparison, multi-methods including the geo-accumulation index (Igeo), absolute principal component score-multiple linear regression model (APCS-MLR), ecological risk index (RI), and sediment quality guidelines (SQGs) were utilized this time, leading to a better result. Findings reveal that pollution levels are generally low or non-existent, with only 1.64% of sampling sites showing close to moderate pollution levels for Cu, Pb, and Zn, and 4.92% and 1.64% of sites exhibiting close to moderate and moderate pollution levels for Cd, respectively. The main contributors to heavy metal presence are pinpointed as industrial wastewater discharge, particularly Cu, Zn, Pb, Cd, and Hg. The ecological risks are also relatively low, with 4.92%, 1.64%, and 1.64% of sampling sites demonstrating close to moderate, moderate, and strong risks in the inflow rivers, respectively. Additionally, only one site shows moderate potential biological toxicity, while the rest display non-toxicity. These findings will update our cognition and offer a scientific basis for pollution treatment and ecosystem enhancement for government management. Full article
(This article belongs to the Special Issue Soil and Groundwater Quality and Resources Assessment)
Show Figures

Figure 1

Figure 1
<p>Location of sampling sites and study area.</p>
Full article ">Figure 2
<p>The <span class="html-italic">Igeo</span> values and pollution levels of heavy metals at sampling sites.</p>
Full article ">Figure 2 Cont.
<p>The <span class="html-italic">Igeo</span> values and pollution levels of heavy metals at sampling sites.</p>
Full article ">Figure 3
<p>Gravel figure of principal component analysis of heavy metals at sampling sites.</p>
Full article ">Figure 4
<p>Statistical figure of <span class="html-italic">E<sub>i</sub></span> and <span class="html-italic">RI</span> for heavy metals at sampling sites.</p>
Full article ">Figure 5
<p>Statistical figure of <span class="html-italic">I<sub>SQGs</sub></span> for heavy metals at sampling sites.</p>
Full article ">Figure 6
<p>Statistical figure of <span class="html-italic">ΣTUs</span> for heavy metals at sampling sites.</p>
Full article ">Figure 7
<p>Statistical figure of the frequency of toxic effects for heavy metals Cu, Pb, Zn, Cr, Ni, Cd, As, and Hg based on <span class="html-italic">C<sub>i</sub></span>, <span class="html-italic">TEC</span>, and <span class="html-italic">PEC</span> at sampling sites.</p>
Full article ">Figure 7 Cont.
<p>Statistical figure of the frequency of toxic effects for heavy metals Cu, Pb, Zn, Cr, Ni, Cd, As, and Hg based on <span class="html-italic">C<sub>i</sub></span>, <span class="html-italic">TEC</span>, and <span class="html-italic">PEC</span> at sampling sites.</p>
Full article ">Figure 7 Cont.
<p>Statistical figure of the frequency of toxic effects for heavy metals Cu, Pb, Zn, Cr, Ni, Cd, As, and Hg based on <span class="html-italic">C<sub>i</sub></span>, <span class="html-italic">TEC</span>, and <span class="html-italic">PEC</span> at sampling sites.</p>
Full article ">
32 pages, 14893 KiB  
Article
Mapping of Clay Montmorillonite Abundance in Agricultural Fields Using Unmixing Methods at Centimeter Scale Hyperspectral Images
by Etienne Ducasse, Karine Adeline, Audrey Hohmann, Véronique Achard, Anne Bourguignon, Gilles Grandjean and Xavier Briottet
Remote Sens. 2024, 16(17), 3211; https://doi.org/10.3390/rs16173211 - 30 Aug 2024
Viewed by 1173
Abstract
The composition of clay minerals in soils, and more particularly the presence of montmorillonite (as part of the smectite family), is a key factor in soil swell–shrinking as well as off–road vehicle mobility. Detecting these topsoil clay minerals and quantifying the montmorillonite abundance [...] Read more.
The composition of clay minerals in soils, and more particularly the presence of montmorillonite (as part of the smectite family), is a key factor in soil swell–shrinking as well as off–road vehicle mobility. Detecting these topsoil clay minerals and quantifying the montmorillonite abundance are a challenge since they are usually intimately mixed with other minerals, soil organic carbon and soil moisture content. Imaging spectroscopy coupled with unmixing methods can address these issues, but the quality of the estimation degrades the coarser the spatial resolution is due to pixel heterogeneity. With the advent of UAV-borne and proximal hyperspectral acquisitions, it is now possible to acquire images at a centimeter scale. Thus, the objective of this paper is to evaluate the accuracy and limitations of unmixing methods to retrieve montmorillonite abundance from very-high-resolution hyperspectral images (1.5 cm) acquired from a camera installed on top of a bucket truck over three different agricultural fields, in Loiret department, France. Two automatic endmember detection methods based on the assumption that materials are linearly mixed, namely the Simplex Identification via Split Augmented Lagrangian (SISAL) and the Minimum Volume Constrained Non-negative Matrix Factorization (MVC-NMF), were tested prior to unmixing. Then, two linear unmixing methods, the fully constrained least square method (FCLS) and the multiple endmember spectral mixture analysis (MESMA), and two nonlinear unmixing ones, the generalized bilinear method (GBM) and the multi-linear model (MLM), were performed on the images. In addition, several spectral preprocessings coupled with these unmixing methods were applied in order to improve the performances. Results showed that our selected automatic endmember detection methods were not suitable in this context. However, unmixing methods with endmembers taken from available spectral libraries performed successfully. The nonlinear method, MLM, without prior spectral preprocessing or with the application of the first Savitzky–Golay derivative, gave the best accuracies for montmorillonite abundance estimation using the USGS library (RMSE between 2.2–13.3% and 1.4–19.7%). Furthermore, a significant impact on the abundance estimations at this scale was in majority due to (i) the high variability of the soil composition, (ii) the soil roughness inducing large variations of the illumination conditions and multiple surface scatterings and (iii) multiple volume scatterings coming from the intimate mixture. Finally, these results offer a new opportunity for mapping expansive soils from imaging spectroscopy at very high spatial resolution. Full article
(This article belongs to the Special Issue Remote Sensing for Geology and Mapping)
Show Figures

Figure 1

Figure 1
<p>Site locations from AGEOTHYP project depicted with colored squares on: (<b>a</b>) topographic map by IGN (National Institute of Geographic and Forest Information) overlaid with smectite abundance from XRD analyses and (<b>b</b>) BRGM swelling hazard map. Soil digital photos of the three selected sites: (<b>c</b>) “Le Buisson” located in Coinces, (<b>d</b>) “Les Laps” located in Gémigny and (<b>e</b>) “La Malandière” located in Mareau.</p>
Full article ">Figure 2
<p>Acquisition setup with the HySpex cameras, RGB composite image from HySpex VNIR camera on Gémigny, Coinces and Mareau sites, with the sampling grid composed of 15 subzones (named after “SUB”), samples collected for laboratory soil characterization in subzones are delimited by red squares (<b>right</b>).</p>
Full article ">Figure 3
<p>NDVI and CAI values for the Mareau hyperspectral image. In red: the thresholds chosen for each index in order to characterize four classes.</p>
Full article ">Figure 4
<p>Grain size and SOC for each site (<b>left</b>), texture triangle for all samples (<b>right</b>).</p>
Full article ">Figure 5
<p>Processing scheme to estimate montmorillonite abundance.</p>
Full article ">Figure 6
<p>Endmembers from laboratory spectral libraries: (<b>a</b>) montmorillonite, (<b>b</b>) kaolinite, (<b>c</b>) illite, (<b>d</b>) quartz and (<b>e</b>) calcite.</p>
Full article ">Figure 7
<p>EM estimates over the Gémigny image. Comparison of the detected and Ducasse EM spectra and graphs of mixture simplex in the first two components space (PC 1 and PC 2) for (<b>a</b>) SISAL to detect 4 EM, (<b>b</b>) SISAL to detect 5 EM, (<b>c</b>) MVC-NMF to detect 4 EM and (<b>d</b>) MVC-NMF to detect 5 EM.</p>
Full article ">Figure 8
<p>Montmorillonite abundance estimations over all the subzones per site (gray boxplots with the median highlighted by a red line) compared to the XRD dataset (boxplots with a red square depicting the median). The inputs are the USGS library, the six preprocessings and REF followed by MLM.</p>
Full article ">Figure 9
<p>Montmorillonite abundance estimations over all the subzones per site (gray boxplots with the median highlighted by a red line) compared to the XRD dataset (boxplots with a red square depicting the median). The inputs are the Ducasse library, the six preprocessings and REF followed by MLM.</p>
Full article ">Figure 10
<p>Performances of Montmorillonite abundance estimations (wt%) obtained with (<b>a</b>) REF-MLM and (<b>b</b>) 1stSGD-MLM with the USGS library (red) and Ducasse spectral library (blue). Bars in the x axis correspond to the accuracy of XRD analysis, and bars in the y axis correspond to the standard deviation of estimated montmorillonite abundances.</p>
Full article ">Figure 11
<p>Results on Gémigny-SUB14: (<b>a</b>) RGB image (in black: masked areas), (<b>b</b>) hillshade map, (<b>c</b>) hillshade histogram (the red vertical line represents the median), (<b>d</b>) difference between the estimated montmorillonite abundance map obtained with REF-MLM and the XRD measured value (in white: masked areas), (<b>g</b>) the same for 1stSGD-MLM, (<b>e</b>) <span class="html-italic">p</span> value maps for REF-MLM (in white: masked areas), (<b>h</b>) the same for 1stSGD-MLM, (<b>f</b>) <span class="html-italic">p</span> value histogram for REF-MLM (the red vertical line represents the median) and (<b>i</b>) the same for 1stSGD-MLM.</p>
Full article ">Figure 12
<p>Results on Coinces-SUB2: (<b>a</b>) RGB image (in black: masked areas), (<b>b</b>) hillshade map, (<b>c</b>) hillshade histogram (the red vertical line represents the median), (<b>d</b>) difference between the estimated montmorillonite abundance map obtained with REF-MLM and the XRD measured value (in white: masked areas), (<b>g</b>) the same for 1stSGD-MLM, (<b>e</b>) <span class="html-italic">p</span> value maps for REF-MLM (in white: masked areas), (<b>h</b>) the same for 1stSGD-MLM, (<b>f</b>) <span class="html-italic">p</span> value histogram for REF-MLM (the red vertical line represents the median) and (<b>i</b>) the same for 1stSGD-MLM.</p>
Full article ">Figure 13
<p>Performances for Montmorillonite abundance estimation with REF-MLM for all subsites (gray boxplots with the median highlighted by a red line) plotted with the XRD dataset (boxplots with a red square depicting the median).</p>
Full article ">Figure 14
<p>Maps for Gémigny site (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 15
<p>Maps for Coinces with wet area SUB10 site (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 16
<p>Maps for Mareau site with wet area SUB15 (<b>a</b>) RGB composite image, (<b>b</b>) composite mask and (<b>c</b>) abundance map of montmorillonite obtained with the REF-MLM and USGS library.</p>
Full article ">Figure 17
<p>Comparison between mineral abundance estimations with REF-MLM and USGS library and the XRD dataset for each site: (<b>a</b>) Coinces, (<b>b</b>) Gémigny, (<b>c</b>) Mareau.</p>
Full article ">
26 pages, 9607 KiB  
Article
A Global Spatial-Spectral Feature Fused Autoencoder for Nonlinear Hyperspectral Unmixing
by Mingle Zhang, Mingyu Yang, Hongyu Xie, Pinliang Yue, Wei Zhang, Qingbin Jiao, Liang Xu and Xin Tan
Remote Sens. 2024, 16(17), 3149; https://doi.org/10.3390/rs16173149 - 26 Aug 2024
Viewed by 882
Abstract
Hyperspectral unmixing (HU) aims to decompose mixed pixels into a set of endmembers and corresponding abundances. Deep learning-based HU methods are currently a hot research topic, but most existing unmixing methods still rely on per-pixel training or employ convolutional neural networks (CNNs), which [...] Read more.
Hyperspectral unmixing (HU) aims to decompose mixed pixels into a set of endmembers and corresponding abundances. Deep learning-based HU methods are currently a hot research topic, but most existing unmixing methods still rely on per-pixel training or employ convolutional neural networks (CNNs), which overlook the non-local correlations of materials and spectral characteristics. Furthermore, current research mainly focuses on linear mixing models, which limits the feature extraction capability of deep encoders and further improvement in unmixing accuracy. In this paper, we propose a nonlinear unmixing network capable of extracting global spatial-spectral features. The network is designed based on an autoencoder architecture, where a dual-stream CNNs is employed in the encoder to separately extract spectral and local spatial information. The extracted features are then fused together to form a more complete representation of the input data. Subsequently, a linear projection-based multi-head self-attention mechanism is applied to capture global contextual information, allowing for comprehensive spatial information extraction while maintaining lightweight computation. To achieve better reconstruction performance, a model-free nonlinear mixing approach is adopted to enhance the model’s universality, with the mixing model learned entirely from the data. Additionally, an initialization method based on endmember bundles is utilized to reduce interference from outliers and noise. Comparative results on real datasets against several state-of-the-art unmixing methods demonstrate the superior of the proposed approach. Full article
(This article belongs to the Topic Computer Vision and Image Processing, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of autoencoder architecture: (<b>a</b>) Autoencoder architecture. (<b>b</b>) Several common decoder architectures.</p>
Full article ">Figure 2
<p>The architecture of the proposed AE network for hyperspectral unmixing.</p>
Full article ">Figure 3
<p>The architecture of the Spatial-Spectral Feature Extraction Module.</p>
Full article ">Figure 4
<p>Module of Multi-Head Self-Attention Modules based on Linear Projection.</p>
Full article ">Figure 5
<p>Dataset: (<b>a</b>) Samson dataset. (<b>b</b>) Jasper Ridge dataset. (<b>c</b>) Urban dataset.</p>
Full article ">Figure 6
<p>The flowchart of the proposed endmember initialization method.</p>
Full article ">Figure 7
<p>The results of endmember extraction (Urban dataset): extracted endmembers (blue) and actual endmembers (orange).</p>
Full article ">Figure 8
<p>Visualization results of endmember bundle extraction (Urban dataset).</p>
Full article ">Figure 9
<p>Abundance maps of tree, water, dirt, and road on the Jasper Ridge dataset obtained by different modules.</p>
Full article ">Figure 10
<p>The results of mRMSE and mSAD under different projection dimensions, along with the corresponding computation times (measured in seconds). (<b>a</b>) Samson dataset. (<b>b</b>) Jasper Ridge dataset.</p>
Full article ">Figure 11
<p>Abundance maps of soil, tree, water on the Samson dataset obtained by different methods.</p>
Full article ">Figure 12
<p>Extracted endmember comparison between the different algorithms and the corresponding GTs in the Samson dataset.</p>
Full article ">Figure 13
<p>Abundance maps of tree, water, dirt, road on the Jasper Ridge dataset obtained by different methods.</p>
Full article ">Figure 14
<p>Extracted endmember comparison between the different algorithms and the corresponding GTs in the Jasper Ridge dataset.</p>
Full article ">Figure 15
<p>Abundance maps of asphalt, grass, tree, roof on the Urban dataset obtained by different methods.</p>
Full article ">Figure 16
<p>Extracted endmember comparison between the different algorithms and the corresponding GTs in the Urban dataset.</p>
Full article ">
21 pages, 5722 KiB  
Article
Seasonal Dynamics of the Land-Surface Characteristics in Arid Regions Retrieved by Optical and Microwave Satellite Data
by Ying Tian, Kurt Ackermann, Christopher McCarthy, Troy Sternberg, Myagmartseren Purevtseren, Che Limuge, Katsuro Hagiwara, Kenta Ogawa, Satoru Hobara and Buho Hoshino
Remote Sens. 2024, 16(17), 3143; https://doi.org/10.3390/rs16173143 - 26 Aug 2024
Viewed by 861
Abstract
Establishing a quantitative relationship between Synthetic Aperture Radar (SAR) data and optical data can facilitate the fusion of these two data sources, enhancing the time-series monitoring capabilities for remote sensing of a land surface. In this study, we analyzed the Normalized Difference Vegetation [...] Read more.
Establishing a quantitative relationship between Synthetic Aperture Radar (SAR) data and optical data can facilitate the fusion of these two data sources, enhancing the time-series monitoring capabilities for remote sensing of a land surface. In this study, we analyzed the Normalized Difference Vegetation Index (NDVI) and Shortwave Infrared Transformed Reflectance (STR) with the backscatter coefficients in vertical polarization VV (σ0VV) and cross polarization VH (σ0VH) across different seasons. We used optical and microwave satellite data spanning from the southern Gobi Desert region to the steppe region in northern Mongolia. The results indicate a relatively high correlation between the NDVI derived from Sentinel-2 and σ0VH (RVH = 0.29, RVH = 0.44, p < 0.001) and a low correlation between the NDVI and σ0VV (RVH = 0.06, RVH = 0.14, p < 0.01) in the Gobi Desert region during summer and fall. STR showed a positive correlation with both σ0VH and σ0VV except in spring, with the highest correlation coefficients observed in summer (RVV = 0.45, RVV = 0.44, p < 0.001). In the steppe region, significant seasonal variations in the NDVI and σ0VH were noted, with a strong positive correlation peaking in summer (RVH = 0.71, p < 0.001) and an inverse correlation with σ0VV except in summer (RVV = −0.43, RVV = −0.34, RVV = −0.13, p < 0.001). Additionally, STR showed a positive correlation with σ0VH and σ0VV in summer (RVH = 0.40, RVV = 0.39, p < 0.001) and fall (RVH = 0.38, RVV = 0.09, p < 0.01), as well as an inverse correlation in spring (RVH= −0.17, RVV= −0.38, p < 0.001) and winter (RVH = −0.21, RVV = −0.06, p < 0.001). The correlations between the NDVI, STR, σ0VH, and σ0VV were shown to vary by season and region. In the Gobi Desert region, perennial shrubs are not photosynthetic in spring and winter, and they affect backscatter due to surface roughness. In the steppe region, annual shrubs were found to be the dominant species and were found to photosynthesize in spring, but not enough to affect the backscatter due to surface roughness. Full article
Show Figures

Figure 1

Figure 1
<p>A schematic of the seasonal dust-outbreak patterns (Where, the arrows in the figure indicate the direction of wind).</p>
Full article ">Figure 2
<p>The study area. Top panel: Mongolian cites of Ulaanbaatar (UB, Tov Province, dark green fill); Mandalgobi (MG, Dundgobi Province, light green fill); Tsogt-Ovoo (TO, Omnogobi Province, brown Fill); and Dalanzadgad (DZ, Omnogobi Province, brown Fill). Bottom panel: (<b>a</b>) elevation; (<b>b</b>) monthly precipitation in August 2019; and (<b>c</b>) the MODIS Maximum NDVI in August 2019 for Gobi Desert sites (1, 2, and 3) and steppe sites (4, 5, and 6).</p>
Full article ">Figure 3
<p>Flowchart of the research methodology. (<b>a</b>) The field investigation contents, including the meteorological station data (precipitation). (<b>b</b>) The remotely sensed data.</p>
Full article ">Figure 4
<p>(<b>a</b>) Land cover distribution map of Mongolia in 2019; and (<b>b</b>) Sentinel−1 Global Backscatter Model of Mongolia (Where, light colors indicate high backscatter coefficients and dark colors indicate low backscatter coefficients. © S1GBM ESA). Reprinted/adapted with permission from Ref. [<a href="#B29-remotesensing-16-03143" class="html-bibr">29</a>].</p>
Full article ">Figure 5
<p>Seasonal variation of the plant height, volume, and soil moisture in the Gobi Desert and steppe sites (the box plots show the mean (square), median (mid−line), first quartile, and third quartile (box edges), as well as the minimum and maximum (whiskers)).</p>
Full article ">Figure 6
<p>Spatial distribution pattern of the vegetation in Gobi Desert and steppe regions.</p>
Full article ">Figure 7
<p>The correlation between the field measurement (2019) of vegetation coverage quadrats (10 m × 10 m) and the unmixed spatial fraction of the vegetation endmember and NDVI calculated from original pixels of Sentinel−2 (2019). (<b>a</b>) The Gobi Desert sites and (<b>b</b>) the steppe sites.</p>
Full article ">Figure 8
<p>(<b>a</b>) The monthly NDVI from Sentinel−2; (<b>b</b>) the monthly STR from Sentinel−2; (<b>c</b>) the monthly σ<sup>0</sup>VH from Sentinel−1; (<b>d</b>) the monthly σ<sup>0</sup>VH from Sentinel−1; and (<b>e</b>) the monthly precipitation (2019).</p>
Full article ">Figure 9
<p>The relationships between the seasonal variation of the NDVI and σ<sup>0</sup>VV and σ<sup>0</sup>VH (dB) in the Gobi Desert and steppe sites. (<b>a</b>) Spring (May); (<b>b</b>) summer (August); (<b>c</b>) fall (November); and (<b>d</b>) winter (January) (Where, the top panel of the figure shows the Gobi Desert sites, and the bottom panel shows the Steppe sites; Also. the red lines are 95% Confidence Interval (CI)).</p>
Full article ">Figure 10
<p>The relationships between the seasonal variation of STR and σ<sup>0</sup>VH and σ<sup>0</sup>VV (dB) in the Gobi Desert and steppe sites. (<b>a</b>) Spring (May); (<b>b</b>) summer (August); (<b>c</b>) fall (November); and (<b>d</b>) winter (January) (Where, the top panel of the figure shows the Gobi Desert sites, and the bottom panel shows the Steppe sites; Also. the red lines are 95% Confidence Interval (CI)).</p>
Full article ">
15 pages, 15270 KiB  
Article
Advanced Unmixing Methodologies for Satellite Thermal Imagery: Matrix Changing and Classification Insights from ASTER and Landsat 8–9
by Paula Andrés-Anaya, Gustavo Hernández-Herráez, Susana Del Pozo and Susana Lagüela
Remote Sens. 2024, 16(16), 3067; https://doi.org/10.3390/rs16163067 - 21 Aug 2024
Viewed by 955
Abstract
The Multisensor Multiresolution Technique (MMT) is applied to unmixed thermal images from ASTER (90 m), using 30 m resolution images from Landsat 8-9 reflective channels. The technique allows for the retrieval of thermal radiance values of the features identified in the high-resolution reflective [...] Read more.
The Multisensor Multiresolution Technique (MMT) is applied to unmixed thermal images from ASTER (90 m), using 30 m resolution images from Landsat 8-9 reflective channels. The technique allows for the retrieval of thermal radiance values of the features identified in the high-resolution reflective images and the generation of a high-resolution radiance image. Different alternatives of application of MMT are evaluated in order to determine the optimal methodology design: performance of the Iterative Self-Organizing Data Analysis Technique (ISODATA) and K-means classification algorithms, with different initiation numbers of clusters, and computation of contributions of each cluster using moving windows with different sizes and with and without weight coefficients. Results show the K-means classification algorithm with five clusters, without matrix weighting, and utilizing a 5 × 5 pixel window for synthetic high-resolution image reconstruction. This approach obtained a maximum R2 of 0.846 and an average R2 of 0.815 across all cases, calculated through the validation of the synthetic high-resolution TIR image generated against a real Landsat 8-9 TIR image from the same area, same date, and co-registered. These values imply a 0.89% improvement regarding the second-best methodology design (K-means with five starting clusters with 7 × 7 moving window) and a 410.25% improvement regarding the worst alternative (K-means with nine initial clusters, weighting, and 3 × 3 moving window). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Workflow of the unmixing methodology proposed.</p>
Full article ">Figure 2
<p>Example of the proposed classification configurations for a chosen study area (Etxalar): (<b>a</b>) ISODATA with 5 initial clusters; (<b>b</b>) K-means with 5 clusters; (<b>c</b>) ISODATA with 9 initial clusters; (<b>d</b>) K-means with 9 clusters.</p>
Full article ">Figure 3
<p>Weight distribution for LR pixels in the computation of the values of the HR pixels within the central LR pixel for cross, 3 × 3, 5 × 5, 7 × 7, and 9 × 9 matrices.</p>
Full article ">Figure 4
<p>Location of the seven case studies selected for the validation of the MMT unmixing optimized methodology. From (<b>left</b>) to (<b>right</b>) and from (<b>top</b>) to (<b>bottom</b>): Genoa, Finland, Marseille, Sicily, Segovia, Etxalar, and Orange.</p>
Full article ">Figure 5
<p>Statistical comparison (R<sup>2</sup>, determination coefficient) of the 30 m synthetic TIR image from ASTER and TIRS using the proposed approaches in the seven case study areas: Segovia, Etxalar, Orange, Marseille, Genova, Ramacca, and Vakka-Suomi. The color ramp used ranges from red, indicating the worst (0.008) adjustment values, to green, indicating the best (0.846), with intermediate values represented as transparent.</p>
Full article ">Figure 6
<p>Statistical comparison (RMSE, root mean square error) of the 30 m synthetic TIR image from ASTER and TIRS using the proposed approaches in the seven case study areas: Segovia, Etxalar, Orange, Marseille, Genova, Ramacca, and Vakka-Suomi. The color ramp used ranges from red, indicating the highest (2.842) error values, to green, indicating the lowest (0.134), with intermediate values represented as transparent.</p>
Full article ">Figure 7
<p>Statistical comparison (MAE, mean absolute error) of the 30 m synthetic TIR image from ASTER and TIRS using the proposed approaches in the seven case study areas: Segovia, Etxalar, Orange, Marseille, Genova, Ramacca, and Vakka-Suomi. The color ramp used ranges from red, indicating the highest (1.023) adjustment values, to green, indicating the lowest (0.086), with intermediate values represented as transparent.</p>
Full article ">Figure 8
<p>Visual comparison of the results of the spatial resolution enhancement of the TOA ASTER TIR radiance data (central image) in W/m<sup>2</sup>sr using Unmixing with K-means classification, 5 × 5 neighborhood, 5 clusters, and unweighted matrix for the case study of Segovia.</p>
Full article ">Figure 9
<p>Visual comparison of the results of the spatial resolution enhancement of the TOA ASTER TIR radiance data (central image) in W/m<sup>2</sup>sr using Unmixing with K-means classification, 5 × 5 neighborhood, 5 clusters, and unweighted matrix for the case study of Etxalar.</p>
Full article ">Figure 10
<p>Visual comparison of the results of the spatial resolution enhancement of the TOA ASTER TIR radiance data (central image) in W/m<sup>2</sup>sr using Unmixing with K-means classification, 5 × 5 neighborhood, 5 clusters, and unweighted matrix for the case study of Orange.</p>
Full article ">
18 pages, 5213 KiB  
Article
Analyzing Thermal Environment Contributions and Driving Factors of LST Heterogeneity in Different Urban Development Zones
by Youshui Zhang, Carlos Alberto Silva and Mengdi Chen
Remote Sens. 2024, 16(16), 2973; https://doi.org/10.3390/rs16162973 - 14 Aug 2024
Cited by 1 | Viewed by 804
Abstract
Analyzing the impacts of urban landscape patterns on the thermal environment has become one of the key research areas in addressing urban heat islands (UHIs) and improving the living environment. A case study was carried out in Fuzhou, Fujian Province of China, and [...] Read more.
Analyzing the impacts of urban landscape patterns on the thermal environment has become one of the key research areas in addressing urban heat islands (UHIs) and improving the living environment. A case study was carried out in Fuzhou, Fujian Province of China, and bi-temporal Landsat imagery was selected to calculate land surface temperature (LST), percent impervious surface area (ISA), and fractional vegetation cover (FVC). The urban area was further divided into three concentric urban zones, ranging from the city center to the urban periphery, based on urban development densities. The spatial pattern of LST and its variance were analyzed and compared between different zones and different dates. The thermal environment contribution of different zones was also quantified to indicate the change in urban landscape patterns resulting from urban expansion in different zones. Furthermore, Geodetector was used to explore the single factors and interaction factors controlling the spatial patterns of LST in each zone. The results showed that (i) urban expansion primarily increased in Zone 2 and Zone 3, and the areal proportion of high and sub-high LST areas increased from 56.11% and 21.08% to 62.03% and 32.49% in Zone 2 and Zone 3, respectively, from 2004 to 2021; (ii) the heat effect contribution of Zones 2 and 3 reached from 75.16% in 2004 to 89.40% in 2021, indicating that the increase in ISA with >LSTmean was more pronounced in Zone 3 and Zone 2 during the period; (iii) the driving factors of LST spatial distribution were regionally different because of the different landscape patterns, and the explanatory power for the heterogeneity of LST in Zone 1 was weaker than in Zone 2 and Zone 3 in the study area; (iv) the interaction of different factors had a higher explanatory power in the spatial distribution of LST than a single factor in each zone because the distributions of land cover types are heterogeneous in urban areas. The results of this study can be used to improve urban planning for urban ecology and UHI mitigation. Full article
Show Figures

Figure 1

Figure 1
<p>The location and Landsat 8 image acquired on 27 September 2021 (a false-color composite image: red, band 5; green, band 4; blue, band 3) of the study area. (The study area was divided into three sectors from the city center to urban peripheral areas in the right figure. The five sample plots delineated with polygons, representing test sites, were used for accuracy assessment in <a href="#remotesensing-16-02973-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 2
<p>Flowchart showing the steps in the study.</p>
Full article ">Figure 3
<p>Spatial distribution patterns of LST and LST classification: (<b>a</b>) Spatial distribution patterns of LST in 2004. (<b>b</b>) Spatial distribution patterns of LST in 2021. (<b>c</b>) Four LST classes in 2004. (<b>d</b>) Four LST classes in 2021.</p>
Full article ">Figure 4
<p>Fractional ISA and FVC derived from ETM+/OLI images using FCLS: (<b>a</b>) 2004 fractional ISA, (<b>b</b>) 2004 FVC, (<b>c</b>) 2021 fractional ISA, and (<b>d</b>) 2021 FVC.</p>
Full article ">Figure 4 Cont.
<p>Fractional ISA and FVC derived from ETM+/OLI images using FCLS: (<b>a</b>) 2004 fractional ISA, (<b>b</b>) 2004 FVC, (<b>c</b>) 2021 fractional ISA, and (<b>d</b>) 2021 FVC.</p>
Full article ">Figure 5
<p>Thermal environment contribution index of each zone to the whole study area of Fuzhou.</p>
Full article ">Figure 6
<p>Thermal environment effect of each zone (WLUI and RWLUI) in the study area of Fuzhou in (<b>a</b>) 2004 and (<b>b</b>) 2021.</p>
Full article ">
22 pages, 7865 KiB  
Article
Applying Knowledge-Based and Data-Driven Methods to Improve Ore Grade Control of Blast Hole Drill Cuttings Using Hyperspectral Imaging
by Somaieh Akbar, Mehdi Abdolmaleki, Saleh Ghadernejad and Kamran Esmaeili
Remote Sens. 2024, 16(15), 2823; https://doi.org/10.3390/rs16152823 - 1 Aug 2024
Cited by 2 | Viewed by 1123
Abstract
This study introduces a novel method utilizing hyperspectral imaging for instantaneous ore-waste analysis of drill cuttings. To implement this technique, we collected samples of drill cuttings at regular depth intervals from five blast holes in an open pit gold mine and subjected them [...] Read more.
This study introduces a novel method utilizing hyperspectral imaging for instantaneous ore-waste analysis of drill cuttings. To implement this technique, we collected samples of drill cuttings at regular depth intervals from five blast holes in an open pit gold mine and subjected them to scanning using a hyperspectral imaging system. Subsequently, we employed two distinct methods for processing the hyperspectral images. A knowledge-based method was used to estimate ore grade within each sampled interval, and a data-driven technique was employed to distinguish the ore and waste for each sample interval. Firstly, leveraging the mixed mineralogical composition of the samples, the Linear Spectral Unmixing (LSU) technique was utilized to predict ore grade for each sample. Additionally, the Gradient Boosting Classifier (GBC) was used as an efficient data-driven approach to classify ore-waste samples. Both methods rendered accurate results when they were compared with results obtained through laboratory X-ray diffraction (XRD) analysis and gold assay analysis for the same sample intervals. Adopting the proposed methodology in open pit mine operations can significantly enhance the process of grade control during blast hole drilling. This includes reducing costs, saving time, minimizing uncertainty in ore grade estimation, and establishing more precise ore-waste boundaries in resource block models. Full article
Show Figures

Figure 1

Figure 1
<p>Location and lithologic map of Gold Bar North.</p>
Full article ">Figure 2
<p>Blast holes: (<b>a</b>) plan view of the blast hole patterns; (<b>b</b>) cross-sections of the five blast holes with the location of Yellowboy fault ore zone; (<b>c</b>) top view of blast holes drilled on a bench; (<b>d</b>) isometric view of blast holes with the pile of drill cuttings.</p>
Full article ">Figure 3
<p>Sampling the drill cuttings every 2 feet (61 cm) along each hole depth. The measured gold assay (ppm) for each interval above the cut-off grade (0.17 ppm) has been shown in blue; for below the cut-off grade, it has been shown in orange.</p>
Full article ">Figure 4
<p>(<b>a</b>) Hyperspectral VNIR RGB image (R: 702 nm G: 548 nm B: 470 nm); (<b>b</b>) hyperspectral SWIR RGB image (R: 1655 nm G: 2200 nm B: 2325 nm) of the drill cutting samples.</p>
Full article ">Figure 5
<p>Workflow for the data-driven and knowledge-based approaches used for ore/waste discrimination.</p>
Full article ">Figure 6
<p>A schematic overview of hyperspectral imagery and Linear Spectral Unmixing technique. Here, m1, m2, and m3 represent different material, while <span class="html-italic">α</span>1, <span class="html-italic">α</span>2, and <span class="html-italic">α</span>3 denote the relative amount of each material.</p>
Full article ">Figure 7
<p>The spectral library was collected over the mine region.</p>
Full article ">Figure 8
<p>Workflow of the classification procedure using.</p>
Full article ">Figure 9
<p>Train and test dataset. Blue parts: train data. Red parts: test data.</p>
Full article ">Figure 10
<p>Mapping endmembers using Linear Spectral Unmixing, Red rectangles indicate selected samples for XRD analysis. The blue bars on the right-hand side show the grade of each tray sample, and the orange bars show the samples with grades below the cut-off grade.</p>
Full article ">Figure 11
<p>Relation between (illite-Calcite + Quartz)% of spectral endmembers and gold assay (ppm). The orange line shows the cut-off grade of gold (0.17 ppm).</p>
Full article ">Figure 12
<p>The visualization of GBC’s performance on the testing dataset. Green: Ore, Orange: Waste, Purple: Misclassified, along with a confusion matrix analysis and accuracy assessment for the test dataset.</p>
Full article ">Figure 13
<p>The degree of importance of the spectral features used in the developed GBC model.</p>
Full article ">Figure 14
<p>The detailed analysis of the frequency of the absorption peak at the most important wavelengths of the four sub-samples taken from the BH 188-10-12. The blue dashed line represents the average absorption peak frequency of waste samples at the given wavelength, and the red dashed line represents the average absorption peak frequency of ore samples at the given wavelength. The green column represents the BH 188-10-12 sub-sample correctly predicted as Ore. The purple columns represent other sub-samples of the BH 188-10-12, which have been misclassified as Waste.</p>
Full article ">
21 pages, 13864 KiB  
Article
A Spectral and Spatial Comparison of Satellite-Based Hyperspectral Data for Geological Mapping
by Rupsa Chakraborty, Imane Rachdi, Samuel Thiele, René Booysen, Moritz Kirsch, Sandra Lorenz, Richard Gloaguen and Imane Sebari
Remote Sens. 2024, 16(12), 2089; https://doi.org/10.3390/rs16122089 - 9 Jun 2024
Cited by 3 | Viewed by 2562
Abstract
The new generation of satellite hyperspectral (HS) sensors provides remarkable potential for regional-scale mineralogical mapping. However, as with any satellite sensor, mapping results are dependent on a typically complex correction procedure needed to remove atmospheric, topographic and geometric distortions before accurate reflectance spectra [...] Read more.
The new generation of satellite hyperspectral (HS) sensors provides remarkable potential for regional-scale mineralogical mapping. However, as with any satellite sensor, mapping results are dependent on a typically complex correction procedure needed to remove atmospheric, topographic and geometric distortions before accurate reflectance spectra can be retrieved. These are typically applied by the satellite operators but use different approaches that can yield different results. In this study, we conduct a comparative analysis of PRISMA, EnMAP, and EMIT hyperspectral satellite data, alongside airborne data acquired by the HyMap sensor, to investigate the consistency between these datasets and their suitability for geological mapping. Two sites in Namibia were selected for this comparison, the Marinkas-Quellen and Epembe carbonatite complexes, based on their geological significance, relatively good exposure, arid climate and data availability. We conducted qualitative and three different quantitative comparisons of the hyperspectral data from these sites. These included correlative comparisons of (1) the reflectance values across the visible-near infrared (VNIR) to shortwave infrared (SWIR) spectral ranges, (2) established spectral indices sensitive to minerals we expect in each of the scenes, and (3) spectral abundances estimated using linear unmixing. The results highlighted a notable shift in inter-sensor consistency between the VNIR and SWIR spectral ranges, with the VNIR range being more similar between the compared sensors than the SWIR. Our qualitative comparisons suggest that the SWIR spectra from the EnMAP and EMIT sensors are the most interpretable (show the most distinct absorption features) but that latent features (i.e., endmember abundances) from the HyMap and PRISMA sensors are consistent with geological variations. We conclude that our results reinforce the need for accurate radiometric and topographic corrections, especially for the SWIR range most commonly used for geological mapping. Full article
(This article belongs to the Topic Advances in Earth Observation and Geosciences)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of the geographical extents covered by the three satellites and one airborne sensor investigated in this study (the EMIT scene is cut for visual purpose).</p>
Full article ">Figure 2
<p>Location (<b>A</b>) and geology of the targeted sites: Epembe (<b>B</b>) (after Geological Survey of Namibia (GSN) and [<a href="#B29-remotesensing-16-02089" class="html-bibr">29</a>]) and Marinkas-Quellen (<b>C</b>) (after [<a href="#B30-remotesensing-16-02089" class="html-bibr">30</a>,<a href="#B34-remotesensing-16-02089" class="html-bibr">34</a>]).</p>
Full article ">Figure 3
<p>False color composites (FCCs) of SWIR bands (R = 2200 nm, G = 2250 nm, and B = 2350 nm) for the available sensors. For Marinkas (left column) (<b>A</b>–<b>D</b>) and Epembe (right column) (<b>E</b>–<b>H</b>) shows FCC from HyMap, EnMap, PRISMA and EMIT, respectively.</p>
Full article ">Figure 4
<p>Two ground sampling points within the calcio-carbonatite lithology at Marinkas-Quellen namely, MQ9 and MQ10 (the colors in the geology map are described in <a href="#remotesensing-16-02089-f002" class="html-fig">Figure 2</a>) (<b>A</b>). At 1 mm spatial sampling (<b>B</b>), MQ9 shows absorptions at 580, 750 and 800 nm, indicating the presence of Nd. The calcite absorption at 2337 nm is present in both MQ9 and MQ10. At 5 m (HyMap), 30 m (EnMAP and PRISMA), and 60 m (EMIT), only the calcite absorption is apparent (<b>C</b>). Note the large jump in estimated reflectance in the PRISMA sensor at 1000 nm, due to the transition from the VNIR to SWIR sensor, and the difficult-to-interpret features in the SWIR range. The spectral graphs are offset for clarity using an offset of ±0.1. The negative values in MQ10 are also due to this offset.</p>
Full article ">Figure 5
<p>Spectral endmembers EM), namely, calcite, dolomite and country rock, selected for Marinkas-Quellen across all four sensors (<b>A</b>–<b>C</b>). The spatial locations of the endmembers are also shown (<b>D</b>). The red dots in (<b>D</b>) marks the locations of the spectra.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>C</b>) show the spectra of specific pixels (blue stars in (<b>D</b>)) from each sensor, highlighting significant variations in the estimated reflectance spectra, especially in the SWIR range (<b>B</b>,<b>C</b>). Please refer to <a href="#remotesensing-16-02089-f002" class="html-fig">Figure 2</a> for the geological map’s legend. The cyan dots in (<b>D</b>) marks locations of the endmembers.</p>
Full article ">Figure 7
<p>Scatterplots showing the absolute (<math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>a</mi> <mi>b</mi> <mi>s</mi> </mrow> </msub> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>) and relative (<math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>p</mi> <mi>r</mi> <mi>o</mi> <mi>p</mi> </mrow> </msub> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>) consistency across the spectral range for Marinkas-Quellen (lower left panels) and Epembe (upper right panels).</p>
Full article ">Figure 8
<p>False color ternary composite showing spectra index results for the Epembe site. The calcite spectral index (red) clearly delineates the calcite-rich carbonatite dyke in each dataset, although its apparent continuity and thickness varies significantly between datasets. Each of the spectral indices have been scaled to range from 1 (black) to 1.4 (saturated). It is to be noted that the HyMap data has masked out vegetation regions (white patches).</p>
Full article ">Figure 9
<p>Scatterplots showing the feature consistency (<math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>f</mi> <mi>e</mi> <mi>a</mi> <mi>t</mi> </mrow> </msub> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>) metrics for the calcite spectral index results at Marinkas-Quellen (lower left panels) and Epembe (upper right panels).</p>
Full article ">Figure 10
<p>Spectral abundance maps from (<b>A</b>) HyMap, (<b>B</b>) EnMAP, (<b>C</b>) EMIT, and (<b>D</b>) PRIMSA for the Marinkas-Quellen site. The x and the y axes in the spectral plots denote wavelength and reflectance, respectively. Each channel of the spectral abundance map is associated with their corresponding colored endmember spectra in the adjacent spectral plot.</p>
Full article ">Figure 11
<p>Scatterplots showing the latent consistency (<math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>l</mi> <mi>a</mi> <mi>t</mi> </mrow> </msub> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>) metrics for the calcite spectral abundance maps at Marinkas-Quellen (lower left) and Epembe (upper right).</p>
Full article ">
18 pages, 11407 KiB  
Article
Estimation of Rice Plant Coverage Using Sentinel-2 Based on UAV-Observed Data
by Yuki Sato, Takeshi Tsuji and Masayuki Matsuoka
Remote Sens. 2024, 16(9), 1628; https://doi.org/10.3390/rs16091628 - 2 May 2024
Cited by 1 | Viewed by 1509
Abstract
Vegetation coverage is a crucial parameter in agriculture, as it offers essential insight into crop growth and health conditions. The spatial resolution of spaceborne sensors is limited, hindering the precise measurement of vegetation coverage. Consequently, fine-resolution ground observation data are indispensable for establishing [...] Read more.
Vegetation coverage is a crucial parameter in agriculture, as it offers essential insight into crop growth and health conditions. The spatial resolution of spaceborne sensors is limited, hindering the precise measurement of vegetation coverage. Consequently, fine-resolution ground observation data are indispensable for establishing correlations between remotely sensed reflectance and plant coverage. We estimated rice plant coverage per pixel using time-series Sentinel-2 Multispectral Instrument (MSI) data, enabling the monitoring of rice growth conditions over a wide area. Coverage was calculated using unmanned aerial vehicle (UAV) data with a spatial resolution of 3 cm with the spectral unmixing method. Coverage maps were generated every 2–3 weeks throughout the rice-growing season. Subsequently, crop growth was estimated at 10 m resolution through multiple linear regression utilizing Sentinel-2 MSI reflectance data and coverage maps. In this process, a geometric registration of MSI and UAV data was conducted to improve their spatial agreement. The coefficients of determination (R2) of the multiple linear regression models were 0.92 and 0.94 for the Level-1C and Level-2A products of Sentinel-2 MSI, respectively. The root mean square errors of estimated rice plant coverage were 10.77% and 9.34%, respectively. This study highlights the promise of satellite time-series models for accurate estimation of rice plant coverage. Full article
(This article belongs to the Special Issue Application of Satellite and UAV Data in Precision Agriculture)
Show Figures

Figure 1

Figure 1
<p>The study field. Labels A to H represent the study plots.</p>
Full article ">Figure 2
<p>Flowchart of this study’s methodology.</p>
Full article ">Figure 3
<p>Geometric registration process of UAV and Sentinel-2 images. (<b>a</b>) Schematic diagram of the geometric registration process; (<b>b</b>) downscaled UAV NIR image; (<b>c</b>) Sentinel-2 band 8 image.</p>
Full article ">Figure 4
<p>Extraction of paddy field pixels: (<b>a</b>) UAV image on 16 June 2023; (<b>b</b>) Sentinel-2 image on 19 June 2023.</p>
Full article ">Figure 5
<p>Five hundred random points on the RGB image. (<b>a</b>) Water endmember on 16 June 2023; (<b>b</b>) rice endmember on 2 September 2023.</p>
Full article ">Figure 6
<p>Normalization results: (<b>a</b>) before normalization, (<b>b</b>) after normalization.</p>
Full article ">Figure 7
<p>Rice plant coverage map based on UAV images.</p>
Full article ">Figure 8
<p>Comparison of UAV images and rice plant coverage maps for Plot A.</p>
Full article ">Figure 9
<p>Seasonal coverage changes in UAV images by plots.</p>
Full article ">Figure 10
<p>Correlation analysis between estimated coverage and correct labels: (<b>a</b>) Sentinel-2 Level-1C product, (<b>b</b>) Sentinel-2 Level-2A product.</p>
Full article ">Figure 11
<p>Rice plant coverage map based on Sentinel-2 images.</p>
Full article ">Figure 11 Cont.
<p>Rice plant coverage map based on Sentinel-2 images.</p>
Full article ">Figure 12
<p>Seasonal changes in coverage in Sentinel-2 images of plots: (<b>a</b>) Sentinel-2 Level-1C product, (<b>b</b>) Sentinel-2 Level-2A product.</p>
Full article ">Figure 13
<p>Seasonal changes in coverage in Sentinel-2 images of Yamada Nishiki rice plants.</p>
Full article ">Figure 14
<p>UAV image taken on July 29. (<b>a</b>) Floating weeds in plot H. (<b>b</b>) Rice plant variability in plot G.</p>
Full article ">
26 pages, 4823 KiB  
Article
Enhancing Urban Above-Ground Vegetation Carbon Density Mapping: An Integrated Approach Incorporating De-Shadowing, Spectral Unmixing, and Machine Learning
by Guangping Qie, Jianneng Ye, Guangxing Wang and Minzi Wang
Forests 2024, 15(3), 480; https://doi.org/10.3390/f15030480 - 4 Mar 2024
Viewed by 1606
Abstract
Accurately mapping urban above-ground vegetation carbon density presents challenges due to fragmented landscapes, mixed pixels, and shadows induced by buildings and mountains. To address these issues, a novel methodological framework is introduced, utilizing a linear spectral unmixing analysis (LSUA) for shadow removal and [...] Read more.
Accurately mapping urban above-ground vegetation carbon density presents challenges due to fragmented landscapes, mixed pixels, and shadows induced by buildings and mountains. To address these issues, a novel methodological framework is introduced, utilizing a linear spectral unmixing analysis (LSUA) for shadow removal and vegetation information extraction from mixed pixels. Parametric and nonparametric models, incorporating LSUA-derived vegetation fraction, are compared, including linear stepwise regression, logistic model-based stepwise regression, k-Nearest Neighbors, Decision Trees, and Random Forests. Applied in Shenzhen, China, the framework integrates Landsat 8, Pleiades 1A & 1B, DEM, and field measurements. Among the key findings, the shadow removal algorithm is effective in mountainous areas, while LSUA-enhanced models improve urban vegetation carbon density mapping, albeit with marginal gains. Integrating kNN and RF with LSUA reduces errors, and Decision Trees, especially when integrated with LSUA, outperform other models. This study underscores the potential of the proposed framework, particularly the integration of Decision Trees with LSUA, for advancing the accuracy of urban vegetation carbon density mapping. Full article
Show Figures

Figure 1

Figure 1
<p>The study area, Shenzhen city, and the spatial arrangement of sample plots within land use and land cover (LULC) categories were determined through the implementation of stratified random sampling to establish the plot sets.</p>
Full article ">Figure 2
<p>Schematic representation of the field plot data collection process for trees (represented by the large square) and shrubs and grass (depicted by three smaller squares).</p>
Full article ">Figure 3
<p>Correlation coefficient analysis examining the relationship between vegetation carbon density and the Landsat 8 images’ original bands. The symbols in the figure denoted as Carbon, B1, B2, …, B9 correspond to vegetation carbon density, Landsat Band 1, Band 2, …, Band 9, respectively.</p>
Full article ">Figure 4
<p>The fractional images were acquired through a mathematical selection method for decomposing mixed pixels utilizing 4 endmembers.</p>
Full article ">Figure 5
<p>Comparison between pre- and post-shadow removal of Landsat 8: (<b>a</b>) Landsat 8 image before shadow removal displayed in natural color; (<b>b</b>) Landsat 8 image after shadow removal displayed in natural color. The red highlighted area is enlarged to illustrate the differences before and after shadow removal.</p>
Full article ">Figure 6
<p>Comparison of the mapping outcomes for vegetation carbon density utilizing LSR, LMSR, kNN, DT, and RF models with and without the incorporation of the vegetation fraction variable derived from LSUA in the modeling process.</p>
Full article ">Figure 7
<p>Optimization of the number of trees utilized in Random Forests (<b>left</b>) and the integration of Random Forests with LSUA (<b>right</b>).</p>
Full article ">
28 pages, 21321 KiB  
Article
The Improved U-STFM: A Deep Learning-Based Nonlinear Spatial-Temporal Fusion Model for Land Surface Temperature Downscaling
by Shanxin Guo, Min Li, Yuanqing Li, Jinsong Chen, Hankui K. Zhang, Luyi Sun, Jingwen Wang, Ruxin Wang and Yan Yang
Remote Sens. 2024, 16(2), 322; https://doi.org/10.3390/rs16020322 - 12 Jan 2024
Cited by 2 | Viewed by 1835
Abstract
The thermal band of a satellite platform enables the measurement of land surface temperature (LST), which captures the spatial-temporal distribution of energy exchange between the Earth and the atmosphere. LST plays a critical role in simulation models, enhancing our understanding of physical and [...] Read more.
The thermal band of a satellite platform enables the measurement of land surface temperature (LST), which captures the spatial-temporal distribution of energy exchange between the Earth and the atmosphere. LST plays a critical role in simulation models, enhancing our understanding of physical and biochemical processes in nature. However, the limitations in swath width and orbit altitude prevent a single sensor from providing LST data with both high spatial and high temporal resolution. To tackle this challenge, the unmixing-based spatiotemporal fusion model (STFM) offers a promising solution by integrating data from multiple sensors. In these models, the surface reflectance is decomposed from coarse pixels to fine pixels using the linear unmixing function combined with fractional coverage. However, when downsizing LST through STFM, the linear mixing hypothesis fails to adequately represent the nonlinear energy mixing process of LST. Additionally, the original weighting function is sensitive to noise, leading to unreliable predictions of the final LST due to small errors in the unmixing function. To overcome these issues, we selected the U-STFM as the baseline model and introduced an updated version called the nonlinear U-STFM. This new model incorporates two deep learning components: the Dynamic Net (DyNet) and the Chang Ratio Net (RatioNet). The utilization of these components enables easy training with a small dataset while maintaining a high generalization capability over time. The MODIS Terra daytime LST products were employed to downscale from 1000 m to 30 m, in comparison with the Landsat7 LST products. Our results demonstrate that the new model surpasses STARFM, ESTARFM, and the original U-STFM in terms of prediction accuracy and anti-noise capability. To further enhance other STFMs, these two deep-learning components can replace the linear unmixing and weighting functions with minor modifications. As a deep learning-based model, it can be pretrained and deployed for online prediction. Full article
(This article belongs to the Special Issue Remote Sensing for Land Surface Temperature and Related Applications)
Show Figures

Figure 1

Figure 1
<p>The study area in Shenzhen and Dongguan within the GBA, China.</p>
Full article ">Figure 2
<p>The problem is the unmixing function. The red region represents the HCRs, and the black square represents the MODIS pixels. The green region in the figure on the left demonstrates the case when HCRs are across multiple MODIS pixels, and the green region in the right figure demonstrates the case when HCRs are only covered by one MODISpixel as the result of making the coverage faction matrix sparser.</p>
Full article ">Figure 3
<p>The problem with the original weighting function. The red region represents the more sensitive region of the error in <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math>; the blue region represents the less sensitive region.</p>
Full article ">Figure 4
<p>The basic idea of the nonlinear U-STFM.</p>
Full article ">Figure 5
<p>Overall workflow of this study.</p>
Full article ">Figure 6
<p>The workflow for training the unmixing model with DyNet.</p>
Full article ">Figure 7
<p>DyNet training process.</p>
Full article ">Figure 8
<p>Data transformation for training the RatioNet.</p>
Full article ">Figure 9
<p>The training process of the RatioNet.</p>
Full article ">Figure 10
<p>Nonlinear U-STFM prediction workflow.</p>
Full article ">Figure 11
<p>The loss value during the training process.</p>
Full article ">Figure 12
<p>The change ratio prediction for each HCR by DyNet: The red cross mark represents the ground truth, and the median value of the multiple predictions by the different beaches was used as the final prediction of the change ratio of each HCR.</p>
Full article ">Figure 13
<p>1:1 plot for predicting LST on 1 November 2000 with different three date pairs (<b>upper</b>) and the final combination prediction (the median value at the pixel level).</p>
Full article ">Figure 14
<p>The final prediction (1 November 2000) based on combining multiple date triplets. (<b>a</b>) the original MODIS LST on 1 November 2000; (<b>b</b>) the prediction of our model; (<b>c</b>) the Landsat LST; (<b>d</b>) the 1:1 plot between our model prediction and the Landsat LST; (<b>e</b>) the RMSE map between our model prediction and the Landsat LST. (1)–(3) are subareas shown in <a href="#remotesensing-16-00322-f015" class="html-fig">Figure 15</a>.</p>
Full article ">Figure 15
<p>Subarea of <a href="#remotesensing-16-00322-f014" class="html-fig">Figure 14</a>.</p>
Full article ">Figure 16
<p>Prediction for 17 September 2001. (<b>a</b>) the original MODIS LST; (<b>b</b>) the prediction of our model; (<b>c</b>) the Landsat LST; (<b>d</b>) the 1:1 plot between our model prediction and the Landsat LST; (<b>e</b>) the RMSE map between our model prediction and the Landsat LST.</p>
Full article ">Figure 17
<p>Comparison of the prediction for 17 September 2001, with or without data for 1 November 2000. Partial cloud coverage marked by the red circle.</p>
Full article ">Figure 18
<p>The 1:1 plot for multiple date predictions.</p>
Full article ">Figure 19
<p>Comparison with U-STFM on 1 November 2000 with multiple HCR setups; (<b>a</b>) the results under 45 HCRs group; (<b>b</b>) the result under 145 HCRs group; (<b>c</b>) the result under 245 HCRs group; (<b>d</b>) the RMSE boxplot under 45, 145 and 245 HCRs group.</p>
Full article ">Figure 20
<p>Comparison with U-STFM on 17 September 2001, with multiple HCR setups. (<b>a</b>) the results under 45 HCRs group; (<b>b</b>) the result under 145 HCRs group; (<b>c</b>) the result under 245 HCRs group; (<b>d</b>) the RMSE boxplot under 45, 145 and 245 HCRs group.</p>
Full article ">Figure 21
<p>Prediction with the different SNRs for 1 November 2000.</p>
Full article ">Figure 22
<p>Boxplot of prediction with the different SNRs.</p>
Full article ">Figure 23
<p>Comparison with the prediction RMSE with STARFM, ESTARFM, and U-STFM.</p>
Full article ">Figure 23 Cont.
<p>Comparison with the prediction RMSE with STARFM, ESTARFM, and U-STFM.</p>
Full article ">Figure 24
<p>The truncation error between the change ratio at HCR level and the pixel level.</p>
Full article ">Figure 25
<p>The theoretical graph of the weighting function.</p>
Full article ">
Back to TopTop