[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (107)

Search Parameters:
Keywords = AVIRIS

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 15074 KiB  
Article
The Standardized Spectroscopic Mixture Model
by Christopher Small and Daniel Sousa
Remote Sens. 2024, 16(20), 3768; https://doi.org/10.3390/rs16203768 - 11 Oct 2024
Viewed by 298
Abstract
The standardized spectral mixture model combines the specificity of a physically based representation of a spectrally mixed pixel with the generality and portability of a spectral index. Earlier studies have used spectrally and geographically diverse collections of broadband and spectroscopic imagery to show [...] Read more.
The standardized spectral mixture model combines the specificity of a physically based representation of a spectrally mixed pixel with the generality and portability of a spectral index. Earlier studies have used spectrally and geographically diverse collections of broadband and spectroscopic imagery to show that the reflectance of the majority of ice-free landscapes on Earth can be represented as linear mixtures of rock and soil substrates (S), photosynthetic vegetation (V) and dark targets (D) composed of shadow and spectrally absorptive/transmissive materials. However, both broadband and spectroscopic studies of the topology of spectral mixing spaces raise questions about the completeness and generality of the Substrate, Vegetation, Dark (SVD) model for imaging spectrometer data. This study uses a spectrally diverse collection of 40 granules from the EMIT imaging spectrometer to verify the generality and stability of the spectroscopic SVD model and characterize the SVD topology and plane of substrates to assess linearity of spectral mixing. New endmembers for soil and non-photosynthetic vegetation (NPV; N) allow the planar SVD model to be extended to a tetrahedral SVDN model to better accommodate the 3D topology of the mixing space. The SVDN model achieves smaller misfit than the SVD, but does so at the expense of implausible fractions beyond [0, 1]. However, a refined spectroscopic SVD model still achieves small (<0.03) RMS misfit, negligible sensitivity to endmember variability and strongly linear scaling over more than an order of magnitude range of spatial resolution. Full article
Show Figures

Figure 1

Figure 1
<p>Index map for EMIT sample sites. Each of the 33 agricultural sites was chosen on the basis of climate, biome, geologic substrate and cropping stage to maximize diversity of vegetation cover density and soil exposure.</p>
Full article ">Figure 2
<p>(<b>A</b>) EMIT mosaic of sites used in this study. Common linear stretch [0, 0.8] for all granules. (<b>B</b>). EMIT mosaic of sites used in this study. Identical to <a href="#remotesensing-16-03768-f002" class="html-fig">Figure 2</a>A, but with a scene-specific 2% linear stretch for each granule.</p>
Full article ">Figure 2 Cont.
<p>(<b>A</b>) EMIT mosaic of sites used in this study. Common linear stretch [0, 0.8] for all granules. (<b>B</b>). EMIT mosaic of sites used in this study. Identical to <a href="#remotesensing-16-03768-f002" class="html-fig">Figure 2</a>A, but with a scene-specific 2% linear stretch for each granule.</p>
Full article ">Figure 3
<p>Spectral mixing space formed by low-order principal components of the EMIT mosaic (<a href="#remotesensing-16-03768-f002" class="html-fig">Figure 2</a>). Orthogonal projections of PCs 1–3 clearly show prominent apexes corresponding to substrate, vegetation and dark reflectances. The outward convexity in PC 3 reveals an additional non-photosynthetic (N) vegetation endmember. Substrate endmember S corresponds to sandy soils, but pure sands have distinct reflectances and form separate mixing trends with the dark endmember. A linear mixture model using the S, V, D and N endmembers projects the mixing space into a tetrahedron bounded by a convex hull of 6 linear mixing trends, excluding sands, cloud and turbid water.</p>
Full article ">Figure 4
<p>Spectral mixing space and joint characterization for the EMIT mosaic. A 3D embedding derived from Uniform Manifold Approximation and Projection (UMAP) reveals two distinct continua for substrates and vegetation surrounded by a constellation of distinct sand and water body reflectances (<b>top</b>). The joint characterization using UMAP and PC projections combines the global structure of the orthogonal PCs with the local structure preserved by UMAP (<b>bottom</b>). Distinct soil and NPV continua increase in reflectance amplitude with PC distinguishing the substrates (PC1) and vegetation (PC2). NPV spans both continua. A single continuum spanning multiple sample sites splits to yield general soil (S1) and NPV (N1) endmembers while many other site-specific soil continua yield endmembers corresponding to spectrally distinct sands shown in <a href="#remotesensing-16-03768-f005" class="html-fig">Figure 5</a>. In contrast to the distinct soil and sand endmembers, all vegetation forms a single continuum spanned by photosynthetic and non-photosynthetic vegetation endmembers.</p>
Full article ">Figure 5
<p>Reflectance spectra from Soil and NPV continua in <a href="#remotesensing-16-03768-f004" class="html-fig">Figure 4</a>. Two distinct NPV continua (N3 and N4) converge to a single continuum that branches (N2) from the soil continuum to a single higher amplitude NPV endmember (N1). The soil continuum extends from the branch point to a single higher amplitude soil endmember (S1). In parallel to this main soil continuum, seven different soil continua (S3–S9) extend to spectrally distinct sand endmembers. Isolated clusters correspond to geographically distinct pure sands in the Negev desert (S2, S10) and Anza-Borrego desert (S11).</p>
Full article ">Figure 6
<p>(<b>A</b>). EMIT SVD composite from SVDN model. Common linear stretch [0, 1] for all. (<b>B</b>). EMIT NPV + misfit composite from SVDN model. Common linear stretch [0, 1] for NPV and [0, 0.05] for misfit.</p>
Full article ">Figure 6 Cont.
<p>(<b>A</b>). EMIT SVD composite from SVDN model. Common linear stretch [0, 1] for all. (<b>B</b>). EMIT NPV + misfit composite from SVDN model. Common linear stretch [0, 1] for NPV and [0, 0.05] for misfit.</p>
Full article ">Figure 7
<p>Endmember fraction spaces for the SVD, SVDN and NVD models. All models are subsets of the same SVDN endmembers, differing only in the inclusion of the S and N endmembers. Comparing the left and center columns, it is apparent that including the N endmember increases the negative fractions for all endmembers. For the SVDN model, RMS misfit diminishes with increasing NPV fraction, but is greatest for spectra with negative NPV fractions. Note much wider ranges for all fraction distributions for the NVD model.</p>
Full article ">Figure 8
<p>SVD versus SVDN model comparison. The Substrate fraction is most sensitive to the addition of the NPV endmember. However, the Vegetation fraction is quite insensitive and the Dark fraction is most sensitive at low fractions. As expected, the SVD model has somewhat higher misfit, although still relatively low at well under 0.04 for the vast majority of spectra in the mosaic. The negligible number of higher misfit spectra are associated with clouds and high albedo sands, which are not represented in either model. The inset covariability matric shows EM correlations on/above the diagonal and Mutual Information (MI) scores below. Note high correlation and MI for S and N.</p>
Full article ">Figure 9
<p>Raw and modeled EMIT spectra with SVD vs. SVDN model misfit space. The NPV-dominant spectra are modeled more accurately with the SVDN than the SVD model. Sands (1,2,3) have higher misfits for both models because neither has a sand endmember. SVD and SVDN models have 90% and 95% (respectively) of spectra with less than 0.03 misfit. Note expanded reflectance scale on example 7.</p>
Full article ">Figure 10
<p>Endmember sensitivity analysis. Three peripheral spectral endmembers (upper left) for S, V and D yield 3<sup>3</sup> = 27 SVD model permutations. Pairwise combinations of each resulting endmember fraction distribution for the EMIT mosaic yield (<sup>27</sup><sub>2</sub>) = 351 model inversion correlations (inset) for each SVD endmember. S and V fraction distribution correlations are &gt; 0.99 but D fractions go as low as 0.98 because differences among S and V endmembers are amplified in D fractions. Standard deviations among model pairs are &lt; 0.05 for each fraction for 99.8% of all 63,692,800 spectra.</p>
Full article ">Figure 11
<p>Linearity of scaling for the SVDN model. The 4 m resolution AVIRIS-3 line was collected the day after the 47 × 60 m resolution EMIT scene. All fractions scale linearly over the order of magnitude difference in resolution. Dispersion is greater for Substrate and NPV fractions. The slight bias of the fractions relative to 1:1 is consistent with the lower solar elevation at time of EMIT collection. Some of the dispersion about the 1:1 lines results from orthographic displacements between images.</p>
Full article ">Figure A1
<p>SVDN fraction spaces for the AVIRIS-3 and EMIT acquisitions compared in <a href="#remotesensing-16-03768-f011" class="html-fig">Figure 11</a>. As expected, the 4.4 m AVIRIS pixel fractions span a wider range than the more spectrally mixed EMIT pixels. As with the EMIT mosaic, S, V and D fractions are well-bounded while NPV shows a significant percentage of negative fractions.</p>
Full article ">
12 pages, 5988 KiB  
Technical Note
The Physiology of Betula glandusa on Two Sunny Summer Days in the Arctic and Linkages with Optical Imagery
by Cameron Proctor, Nam Leu and Bin Wang
Remote Sens. 2024, 16(12), 2160; https://doi.org/10.3390/rs16122160 - 14 Jun 2024
Viewed by 637
Abstract
Controls on Arctic vegetation physiology have been linked to microscale (1–100 m) topography and landscape position, yet drivers may change under future climates as temperature, active-layer thickness, and nutrient limitations are removed or altered. Focusing on the cosmopolitan dwarf birch (Betula glandusa [...] Read more.
Controls on Arctic vegetation physiology have been linked to microscale (1–100 m) topography and landscape position, yet drivers may change under future climates as temperature, active-layer thickness, and nutrient limitations are removed or altered. Focusing on the cosmopolitan dwarf birch (Betula glandusa), physiological metrics were measured over two field campaigns at Trail Valley Creek, NWT, Canada, and linked to tasked and archived multispectral imagery to investigate drivers. Relative humidity was ~31.1% on 25 June 2023, and increased to 45.6% on 29 June 2023, which corresponded to heightened physiological activity of stomatal conductance and light-adapted fluorescence (gsm: 0.118 vs. 0.165 μmol m−2 s−1, Fs: 129.29 vs. 178.42). Normalized difference vegetation index of AVIRIS, Sentinel 2, and SkySat were negligibly correlated to dwarf birch physiological activity, but moderately correlated to dwarf birch height and active-layer thickness. Random forest variable importance revealed that environmental factors and field-measured active-layer thickness ranked higher than remote sensing metrics in explaining physiological activity regardless of the field campaign. Overall, these findings suggest that microscale variation can influence dwarf birch physiological activity, yet microscale effects are overwritten by environmental conditions that may hinder fine-scale space-based monitoring of Arctic vegetation physiological dynamics. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Trail Valley Creek, NWT is located 50 km north of Inuvik.</p>
Full article ">Figure 2
<p>Dwarf birch along a 100 m long transect from open tundra (left) to a dwarf birch dominant patch (right): (<b>A</b>) Example of a tundra site underlain with lichen and small crown diameter birch. (<b>B</b>) Transition between tundra and a dwarf birch patch with a more enclosed overstory comprised of birch mixed with Labrador tea. (<b>C</b>) Dwarf birch patch with nearly enclosed overstory dominated by dwarf birch with minimal other species in the understory.</p>
Full article ">Figure 3
<p>Trail Valley Creek study area from SkySat tacking including the sample location of the Grid and Circle campaigns. Number indicates site number. <b>Left</b>: true color. <b>Right</b>: NDVI.</p>
Full article ">Figure 4
<p>Dynamics of air temperature (°C) and humidity (%) during the two field data collection campaigns.</p>
Full article ">Figure 5
<p>Stomatal conductance by campaign over time by birch density class. Orange lines represents linear regression line of best fit.</p>
Full article ">Figure 6
<p>Physiological activity of birch as a function of birch density. Data for each campaign displayed with the left exclusive to the Grid data points, and right the Circle data points. Boxplots with different letters are significantly different according to Tukey’s HSD. <b>Top left</b>: stomatal conductance. <b>Top Right</b>: steady-state fluorescence. <b>Bottom Left</b>: maximum fluorescence. <b>Bottom Right</b>: Quantum efficiency of the PSII system.</p>
Full article ">Figure 7
<p>Random forest regression variable importance for the four physiological metrics separated by Gid + Circle (<b>top</b>), Grid only (<b>middle</b>), and Circle only (<b>bottom</b>). Dots are sized by importance and the color denotes whether the coefficient was positive or negative.</p>
Full article ">
17 pages, 2881 KiB  
Article
Expanded Signal to Noise Ratio Estimates for Validating Next-Generation Satellite Sensors in Oceanic, Coastal, and Inland Waters
by Raphael M. Kudela, Stanford B. Hooker, Liane S. Guild, Henry F. Houskeeper and Niky Taylor
Remote Sens. 2024, 16(7), 1238; https://doi.org/10.3390/rs16071238 - 31 Mar 2024
Viewed by 1232
Abstract
The launch of the NASA Plankton, Aerosol, Cloud, ocean Ecosystem (PACE) and the Surface Biology and Geology (SBG) satellite sensors will provide increased spectral resolution compared to existing platforms. These new sensors will require robust calibration and validation datasets, but existing field-based instrumentation [...] Read more.
The launch of the NASA Plankton, Aerosol, Cloud, ocean Ecosystem (PACE) and the Surface Biology and Geology (SBG) satellite sensors will provide increased spectral resolution compared to existing platforms. These new sensors will require robust calibration and validation datasets, but existing field-based instrumentation is limited in its availability and potential for geographic coverage, particularly for coastal and inland waters, where optical complexity is substantially greater than in the open ocean. The minimum signal-to-noise ratio (SNR) is an important metric for assessing the reliability of derived biogeochemical products and their subsequent use as proxies, such as for biomass, in aquatic systems. The SNR can provide insight into whether legacy sensors can be used for algorithm development as well as calibration and validation activities for next-generation platforms. We extend our previous evaluation of SNR and associated uncertainties for representative coastal and inland targets to include the imaging sensors PRISM and AVIRIS-NG, the airborne-deployed C-AIR radiometers, and the shipboard HydroRad and HyperSAS radiometers, which were not included in the original analysis. Nearly all the assessed hyperspectral sensors fail to meet proposed criteria for SNR or uncertainty in remote sensing reflectance (Rrs) for some part of the spectrum, with the most common failures (>20% uncertainty) below 400 nm, but all the sensors were below the proposed 17.5% uncertainty for derived chlorophyll-a. Instrument suites for both in-water and airborne platforms that are capable of exceeding all the proposed thresholds for SNR and Rrs uncertainty are commercially available. Thus, there is a straightforward path to obtaining calibration and validation data for current and next-generation sensors, but the availability of suitable high spectral resolution sensors is limited. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Data collection locations used in this analysis, plotted as solid black circles. (<b>A</b>) Coastal California and the Gulf of Mexico; (<b>B</b>) expanded view for California; (<b>C</b>) expanded view for the Gulf of Mexico; and (<b>D</b>) a dark target from Hawaii, with Marine Optical Buoy (MOBY) for reference.</p>
Full article ">Figure 2
<p>SNR calculated from field measurements for C-AERO (circles) and AVIRIS-C (diamonds). Filled symbols are dark target results; open symbols are bright target results. The solid horizontal lines denote recommended uncertainty levels (values should be below the lines). Reproduced from [<a href="#B21-remotesensing-16-01238" class="html-bibr">21</a>].</p>
Full article ">Figure 3
<p>SNR for dark and bright targets. (<b>A</b>) SNR for Santa Barbara Channel from AVIRIS-NG; (<b>B</b>) SNR for bright targets in Monterey Bay from C-AIR; (<b>C</b>) SNR from the S-MODE site (dark targets) from C-AIR. S-MODE-1 and S-MODE-2 are data flying into and out of the principal plane of the Sun, respectively; (<b>D</b>) SNR for Moss Landing Harbor (bright target) and Hawaii (dark target) from PRISM; (<b>E</b>) SNR for San Francisco Bay and Gulf of Mexico (both bright targets) from HydroRad-3 and HyperSAS. For each panel the proposed SNR thresholds (2, 20–50, 40–100; [<a href="#B21-remotesensing-16-01238" class="html-bibr">21</a>]) are indicated for reference.</p>
Full article ">Figure 4
<p>Data from (<b>A</b>) AVIRIS-NG (Santa Barbara Channel; black = dark target, gray = bright (kelp) target) for <span class="html-italic">R</span><sub>rs</sub> (solid and dashed lines) and <span class="html-italic">R</span><sub>rs</sub> uncertainty; (<b>B</b>) HyperSAS (black; Gulf of Mexico) and HydroRad-3 (gray; San Francisco Bay) for <span class="html-italic">R</span><sub>rs</sub> (black and gray lines) and <span class="html-italic">R</span><sub>rs</sub> uncertainty; (<b>C</b>) PRISM (Hawaii = black, dark target; Moss Landing Harbor = gray, bright target) and <span class="html-italic">R</span><sub>rs</sub> uncertainty. Panels (<b>D</b>–<b>F</b>) show corresponding percent uncertainty with proposed thresholds as solid horizontal black lines (data should be below the line). Missing data at some wavelengths represent failure to converge on an acceptable fit for the variogram analysis. Panel (<b>B</b>) includes a dashed horizontal line for 0 <span class="html-italic">R</span><sub>rs</sub> and <span class="html-italic">R</span><sub>rs</sub> uncertainty; values below the dashed line represent negative calculated reflectances. Data from AVIRIS-NG (panels (<b>A</b>,<b>D</b>)) were truncated at 900 nm.</p>
Full article ">
20 pages, 3969 KiB  
Article
Adapting Prediction Models to Bare Soil Fractional Cover for Extending Topsoil Clay Content Mapping Based on AVIRIS-NG Hyperspectral Data
by Elizabeth Baby George, Cécile Gomez and Nagesh D. Kumar
Remote Sens. 2024, 16(6), 1066; https://doi.org/10.3390/rs16061066 - 18 Mar 2024
Viewed by 1053
Abstract
The deployment of remote sensing platforms has facilitated the mapping of soil properties to a great extent. However, the accuracy of these soil property estimates is compromised by the presence of non-soil cover, which introduces interference with the acquired reflectance spectra over pixels. [...] Read more.
The deployment of remote sensing platforms has facilitated the mapping of soil properties to a great extent. However, the accuracy of these soil property estimates is compromised by the presence of non-soil cover, which introduces interference with the acquired reflectance spectra over pixels. Therefore, current soil property estimation by remote sensing is limited to bare soil pixels, which are identified based on spectral indices of vegetation. Our study proposes a composite mapping approach to extend the soil properties mapping beyond bare soil pixels, associated with an uncertainty map. The proposed approach first classified the pixels based on their bare soil fractional cover by spectral unmixing. Then, a specific regression model was built and applied to each bare soil fractional cover class to estimate clay content. Finally, the clay content maps created for each bare soil fractional cover class were mosaicked to create a composite map of clay content estimations. A bootstrap procedure was used to estimate the standard deviation of clay content predictions per bare soil fractional cover dataset, which represented the uncertainty of estimations. This study used a hyperspectral image acquired by the Airborne Visible/Infrared Imaging Spectrometer-Next Generation (AVIRIS-NG) sensor over cultivated fields in South India. The proposed approach provided modest performances in prediction (Rval2 ranging from 0.53 to 0.63) depending on the bare soil fractional cover class and showed a correct spatial pattern, regardless of the bare soil fraction classes. The model’s performance was observed to increase with the adoption of higher bare soil fractional cover thresholds. The mapped area ranged from 10.4% for pixels with bare soil fractional cover >0.7 to 52.7% for pixels with bare soil fractional cover >0.3. The approach thus extended the mapped surface by 42.4%, while maintaining acceptable prediction performances. Finally, the proposed approach could be adopted to extend the mapping capability of planned and current hyperspectral satellite missions. Full article
(This article belongs to the Special Issue Recent Advances in Remote Sensing of Soil Science)
Show Figures

Figure 1

Figure 1
<p>Location of the study area in (<b>a</b>) the state of Karnataka in India (pink polygon) and (<b>b</b>) Chamarajanagar district (bleu polygon) and (<b>c</b>) locations of the 272 topsoil samples (yellow dots) over the AVIRIS-NG False Colo r Composite (R-G-B as 782 nm, 662 nm, and 552 nm) with a zoom area near Berambadi lake (blue rectangle).</p>
Full article ">Figure 2
<p>The workflow of the methodology with (<b>a</b>) generation of <span class="html-italic">SFClassMap</span>, (<b>b</b>) building the regression models <span class="html-italic">M<sub>p</sub></span> and <span class="html-italic">M<sub>all_p</sub></span>, (<b>c</b>) application of the models on the AVIRIS-NG image, and (<b>d</b>) creation of the <span class="html-italic">CompositeClayMap</span> and <span class="html-italic">UncertaintyMap (</span>“SD” stands for standard deviation).</p>
Full article ">Figure 3
<p>(<b>a</b>) Bare soil fractional cover classes map (<span class="html-italic">SFClassMap</span>), (<b>b</b>) composite clay content map (<span class="html-italic">CompositeClayMap</span>; mean clay content from 100 iterations), and (<b>c</b>) the associated map of uncertainties (<span class="html-italic">UncertaintyMap</span>; standard deviation from 100 iterations) of a sub-part of the study area (blue rectangle in <a href="#remotesensing-16-01066-f001" class="html-fig">Figure 1</a>c). The pink and black ellipses highlight areas characterized by red and black soils, respectively. The purple ellipse highlights an area with high uncertainty.</p>
Full article ">Figure 4
<p>Standard deviation of clay content estimations (from 100 iterations) in the bootstrap procedure. The thin line, bar, and white dot show the 95% confidence interval, interquartile range, and median, respectively.</p>
Full article ">
19 pages, 11920 KiB  
Article
Mapping Leaf Area Index at Various Rice Growth Stages in Southern India Using Airborne Hyperspectral Remote Sensing
by Mathyam Prabhakar, Kodigal A. Gopinath, Nakka Ravi Kumar, Merugu Thirupathi, Uppu Sai Sravan, Golla Srasvan Kumar, Gutti Samba Siva, Pebbeti Chandana and Vinod Kumar Singh
Remote Sens. 2024, 16(6), 954; https://doi.org/10.3390/rs16060954 - 8 Mar 2024
Cited by 2 | Viewed by 1723
Abstract
Globally, rice is one of the most important staple food crops. The most significant metric for evaluating the rice growth and productivity is the Leaf Area Index (LAI), which can be effectively monitored using remote sensing data. Hyperspectral remote sensing provides contiguous bands [...] Read more.
Globally, rice is one of the most important staple food crops. The most significant metric for evaluating the rice growth and productivity is the Leaf Area Index (LAI), which can be effectively monitored using remote sensing data. Hyperspectral remote sensing provides contiguous bands at narrow wavelengths for mapping LAI at various rice phenological stages, and it is functionally related to canopy spectral reflectance. Hyperspectral signatures for different phases of rice crop growth was recorded using Airborne Visible Near-Infrared Imaging Spectrometer-Next Generation (AVIRIS-NG) along with corresponding ground based observations. Ground-based hyperspectral canopy spectral reflectance measurements were recorded with FieldSpec 3 Hi-Res spectroradiometer (ASD Inc., Forsyth County, GA, USA; spectral range: 350–2500 nm) and LAI data from 132 farmer’s fields in Southern India. Among 29 hyperspectral vegetation indices tested, 8 were found promising for mapping rice LAI at various phenological stages. Among all the growth stages, the elongation stage was the most accurately estimated using vegetation indices that exhibited a significant correlation with the airborne hyperspectral reflectance. The validation of hyperspectral vegetation indices revealed that the best fit model for estimating rice LAI was mND705 (red-edge, blue, and NIR bands) at seedling and elongation, SAVI (red and NIR bands) at tillering and WDRVI (red and NIR bands) at booting stage. Full article
(This article belongs to the Special Issue Application of Satellite and UAV Data in Precision Agriculture)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>AVIRIS-NG map with six scenes used for the study (Digitization footprint: 1.85 MB/ha).</p>
Full article ">Figure 2
<p>FCC of the study area and the sampling sites with different color codes (Red: 817 nm, Green: 642 nm, and Blue: 432 nm).</p>
Full article ">Figure 3
<p>Rice crop at various growth phases: (<b>a</b>) seedling; (<b>b</b>) tillering; (<b>c</b>) elongation; (<b>d</b>) heading; (<b>e</b>) flowering; (<b>f</b>) maturity.</p>
Full article ">Figure 4
<p>Hyperspectral signatures for different crop growth stages of rice recorded using (<b>a</b>) ground based measurements and (<b>b</b>) AVIRIS-NG.</p>
Full article ">Figure 5
<p>Mean of vegetation indices at different crop growth stages of rice estimated from AVIRIS-NG. Error bars represents standard deviation. The mean values in the stacked bar for SR, TVI, RVI, and REP are in the range of 60–750, and thus are not shown in the above Figure (Refer to <a href="#remotesensing-16-00954-t0A1" class="html-table">Table A1</a> in <a href="#app2-remotesensing-16-00954" class="html-app">Appendix A</a> for more details).</p>
Full article ">Figure 6
<p>Best fit models for estimating LAI using various Vegetation Indices (VI) at different crop growth stages of rice. Error bars represents RMSE values.</p>
Full article ">Figure 7
<p>LAI of rice fields used for the calibration and validation of data.</p>
Full article ">Figure 8
<p>Estimated Leaf Area Index (LAI) of rice fields with (<b>a</b>) mND<sub>705</sub>; (<b>b</b>) SAVI; (<b>c</b>) WDRVI.</p>
Full article ">Figure A1
<p>Map of India representing the study area.</p>
Full article ">
17 pages, 9275 KiB  
Article
Mapping Soil Organic Carbon Stock Using Hyperspectral Remote Sensing: A Case Study in the Sele River Plain in Southern Italy
by Nicolas Francos, Paolo Nasta, Carolina Allocca, Benedetto Sica, Caterina Mazzitelli, Ugo Lazzaro, Guido D’Urso, Oscar Rosario Belfiore, Mariano Crimaldi, Fabrizio Sarghini, Eyal Ben-Dor and Nunzio Romano
Remote Sens. 2024, 16(5), 897; https://doi.org/10.3390/rs16050897 - 3 Mar 2024
Cited by 2 | Viewed by 3027
Abstract
Mapping soil organic carbon (SOC) stock can serve as a resilience indicator for climate change. As part of the carbon dioxide (CO2) sink, soil has recently become an integral part of the global carbon agenda to mitigate climate change. We used [...] Read more.
Mapping soil organic carbon (SOC) stock can serve as a resilience indicator for climate change. As part of the carbon dioxide (CO2) sink, soil has recently become an integral part of the global carbon agenda to mitigate climate change. We used hyperspectral remote sensing to model the SOC stock in the Sele River plain located in the Campania region in southern Italy. To this end, a soil spectral library (SSL) for the Campania region was combined with an aerial hyperspectral image acquired with the AVIRIS–NG sensor mounted on a Twin Otter aircraft at an altitude of 1433 m. The products of this study were four raster layers with a high spatial resolution (1 m), representing the SOC stocks and three other related soil attributes: SOC content, clay content, and bulk density (BD). We found that the clay minerals’ spectral absorption at 2200 nm has a significant impact on predicting the examined soil attributes. The predictions were performed by using AVIRIS–NG sensor data over a selected plot and generating a quantitative map which was validated with in situ observations showing high accuracies in the ground-truth stage (OC stocks [RPIQ = 2.19, R2 = 0.72, RMSE = 0.07]; OC content [RPIQ = 2.27, R2 = 0.80, RMSE = 1.78]; clay content [RPIQ = 1.6 R2 = 0.89, RMSE = 25.42]; bulk density [RPIQ = 1.97, R2 = 0.84, RMSE = 0.08]). The results demonstrated the potential of combining SSLs with remote sensing data of high spectral/spatial resolution to estimate soil attributes, including SOC stocks. Full article
(This article belongs to the Special Issue Remote Sensing of Carbon Fluxes and Stocks II)
Show Figures

Figure 1

Figure 1
<p>The geographical locations of the Campania SSL samples used for this study.</p>
Full article ">Figure 2
<p>The Campania region provinces and the locations of the Campania SSL samples in which the coordinates were recorded.</p>
Full article ">Figure 3
<p>Positions of soil and non-soil ground-truth targets.</p>
Full article ">Figure 4
<p>Correlation matrix (Pearson’s r) between the examined soil properties in different scenarios: (<b>a</b>) no preprocessing; (<b>b</b>) Log(x); (<b>c</b>) x<sup>2</sup>; (<b>d</b>) 1/x.</p>
Full article ">Figure 5
<p>Certainty matrix (<span class="html-italic">p</span>-values) between the examined soil properties in different scenarios: (<b>a</b>) no preprocessing; (<b>b</b>) Log(x); (<b>c</b>) x<sup>2</sup>; (<b>d</b>) 1/x.</p>
Full article ">Figure 6
<p>Comparison of spectral ground-truth target measurements in the laboratory, field, and by AVIRIS–NG.</p>
Full article ">Figure 7
<p>Spectral-based models: (<b>a</b>) SOC stock; (<b>b</b>) SOC content; (<b>c</b>) clay content; and (<b>d</b>) BD.</p>
Full article ">Figure 8
<p>AVIRIS–NG prediction maps: SOC stock; SOC content; clay content; and BD.</p>
Full article ">
20 pages, 8858 KiB  
Article
Integrating SAR and Optical Data for Aboveground Biomass Estimation of Coastal Wetlands Using Machine Learning: Multi-Scale Approach
by Mohammadali Hemati, Masoud Mahdianpari, Hodjat Shiri and Fariba Mohammadimanesh
Remote Sens. 2024, 16(5), 831; https://doi.org/10.3390/rs16050831 - 28 Feb 2024
Cited by 5 | Viewed by 3387
Abstract
Coastal wetlands encompass diverse ecosystems such as tidal marshes, mangroves, and seagrasses, which harbor substantial amounts of carbon (C) within their vegetation and soils. Despite their relatively small global extent, these wetlands exhibit carbon sequestration rates on par with those observed in terrestrial [...] Read more.
Coastal wetlands encompass diverse ecosystems such as tidal marshes, mangroves, and seagrasses, which harbor substantial amounts of carbon (C) within their vegetation and soils. Despite their relatively small global extent, these wetlands exhibit carbon sequestration rates on par with those observed in terrestrial forests. The application of remote sensing technologies offers a promising means of monitoring aboveground biomass (AGB) in wetland environments. However, the scarcity of field data poses a significant challenge to the utilization of spaceborne data for accurate estimation of AGB in coastal wetlands. To address this limitation, this study presents a novel multi-scale approach that integrates field data, aerial imaging, and satellite platforms to generate high-quality biomass maps across varying scales. At the fine scale level, the AVIRIS-NG hyperspectral data were employed to develop a model for estimating AGB with an exceptional spatial resolution of 5 m. Subsequently, at a broader scale, large-scale and multitemporal models were constructed using spaceborne Sentinel-1 and Sentinel-2 data collected in 2021. The Random Forest (RF) algorithm was utilized to train spring, fall and multi-temporal models using 70% of the available reference data. Using the remaining 30% of untouched data for model validation, Root Mean Square Errors (RMSE) of 0.97, 0.98, and 1.61 Mg ha−1 was achieved for the spring, fall, and multi-temporal models, respectively. The highest R-squared value of 0.65 was achieved for the multi-temporal model. Additionally, the analysis highlighted the importance of various features in biomass estimation, indicating the contribution of different bands and indices. By leveraging the wetland inventory classification map, a comprehensive temporal analysis was conducted to examine the average and total AGB dynamics across various wetland classes. This analysis elucidated the patterns and fluctuations in AGB over time, providing valuable insights into the temporal dynamics of these wetland ecosystems. Full article
(This article belongs to the Special Issue Earth Observation Data in Environmental Data Spaces)
Show Figures

Figure 1

Figure 1
<p>An overview of the study area.</p>
Full article ">Figure 2
<p>Different earth observation data are used for AGB estimation.</p>
Full article ">Figure 3
<p>Flowchart of the proposed method. Boxes are color-coded with orange for spaceborne, blue for airborne, and red for ground information. Yellow boxes denote RF models, green indicates validation, and purple signifies results.</p>
Full article ">Figure 4
<p>Scatter plot of the predicted and actual AGB (Mg ha<sup>−1</sup>) values for: (<b>A</b>) spring, (<b>B</b>) fall, and (<b>C</b>) multi-temporal models. Fitted lines with 95% interval indicators are shown in blue lines.</p>
Full article ">Figure 5
<p>The importance of different features from the RF models.</p>
Full article ">Figure 6
<p>AGB map of the spring season (March and April 2021) and zoomed maps of (<b>A</b>–<b>C</b>) regions. The first row shows results from AVIRIS-NG (small-scale model), and the second row shows results from Sentinel-1 and -2 (large-scale model).</p>
Full article ">Figure 7
<p>AGB map of the fall season (August 2021) and zoomed maps of (<b>A</b>–<b>C</b>) regions. The first row shows results from AVIRIS-NG (small-scale model), and the second row shows results from Sentinel-1 and -2 (large-scale model).</p>
Full article ">Figure 8
<p>AGB monthly maps produced using the multi-temporal model.</p>
Full article ">Figure 9
<p>Average AGB (lines) and estimated total AGB (bars) for each wetland class in the region.</p>
Full article ">
18 pages, 4775 KiB  
Article
A Hyperspectral Change Detection (HCD-Net) Framework Based on Double Stream Convolutional Neural Networks and an Attention Module
by Seyd Teymoor Seydi, Mahboubeh Boueshagh, Foad Namjoo, Seyed Mohammad Minouei, Zahir Nikraftar and Meisam Amani
Remote Sens. 2024, 16(5), 827; https://doi.org/10.3390/rs16050827 - 28 Feb 2024
Cited by 3 | Viewed by 1555
Abstract
Human activities and natural phenomena continually transform the Earth’s surface, presenting ongoing challenges to the environment. Therefore, the accurate and timely monitoring and prediction of these alterations are essential for devising effective solutions and mitigating environmental impacts in advance. This study introduces a [...] Read more.
Human activities and natural phenomena continually transform the Earth’s surface, presenting ongoing challenges to the environment. Therefore, the accurate and timely monitoring and prediction of these alterations are essential for devising effective solutions and mitigating environmental impacts in advance. This study introduces a novel framework, called HCD-Net, for detecting changes using bi-temporal hyperspectral images. HCD-Net is built upon a dual-stream deep feature extraction process, complemented by an attention mechanism. The first stream employs 3D convolution layers and 3D Squeeze-and-Excitation (SE) blocks to extract deep features, while the second stream utilizes 2D convolution and 2D SE blocks for the same purpose. The deep features from both streams are then concatenated and processed through dense layers for decision-making. The performance of HCD-Net is evaluated against existing state-of-the-art change detection methods. For this purpose, the bi-temporal Airborne Visible/Infrared Imaging Spectrometer (AVIRIS) hyperspectral dataset was utilized to assess the change detection performance. The findings indicate that HCD-Net achieves superior accuracy and the lowest false alarm rate among the compared methods, with an overall classification accuracy exceeding 96%, and a kappa coefficient greater than 0.9. Full article
(This article belongs to the Section Remote Sensing Image Processing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of the proposed HCD-Net.</p>
Full article ">Figure 2
<p>The architecture of the proposed CNN framework for HCD.</p>
Full article ">Figure 3
<p>The main difference between 2D/3D Squeeze-and-Excitation (SE) blocks.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) show composite images of the original HSI collected from the China dataset in 2006 and 2007, respectively, and (<b>c</b>) presents a ground truth image in binary format for these datasets. (<b>d</b>,<b>e</b>) show false-color composites of the original hyperspectral images from the USA dataset, acquired in 2004 and 2007, respectively, and (<b>f</b>) is the ground control data in binary format. (<b>g</b>,<b>h</b>) display hyperspectral images captured from the Bay Area dataset in 2013 and 2015, respectively, and (<b>i</b>) presents the ground truth data.</p>
Full article ">Figure 5
<p>Image pixels used for training, validation, and testing in the (<b>a</b>) Bay Area dataset, (<b>b</b>) China dataset, and (<b>c</b>) USA dataset. Training and validation sets are represented by green for No-Change pixels and red for Change pixels, while beige indicates the test set.</p>
Full article ">Figure 6
<p>The output of the binary HCD for the Bay Area dataset. (<b>a</b>) IR-MAD SVM, (<b>b</b>) 2D-Siamese, (<b>c</b>) 3D-Siamese, (<b>d</b>) GETNET, (<b>e</b>) HCD-Net, and (<b>f</b>) Binary Ground Control.</p>
Full article ">Figure 7
<p>The result of the binary HCD for the China dataset. (<b>a</b>) IR-MAD SVM, (<b>b</b>) 2D-Siamese, (<b>c</b>) 3D-Siamese, (<b>d</b>) GETNET, (<b>e</b>) HCD-Net, and (<b>f</b>) Binary Ground Control.</p>
Full article ">Figure 8
<p>The result of the binary HCD for the USA dataset. (<b>a</b>) IR-MAD SVM, (<b>b</b>) 2D-Siamese, (<b>c</b>) 3D-Siamese, (<b>d</b>) GETNET, (<b>e</b>) HCD-Net, and (<b>f</b>) Binary Ground Control.</p>
Full article ">
14 pages, 4054 KiB  
Article
Unsupervised Wavelet-Feature Correlation Ratio Markov Clustering Algorithm for Remotely Sensed Images
by Zhaohui Wang
Appl. Sci. 2024, 14(2), 767; https://doi.org/10.3390/app14020767 - 16 Jan 2024
Viewed by 902
Abstract
The spectrums of one type of object under different conditions have the same features (up, down, protruding, concave) at the same spectral positions, which can be used as primary parameters to evaluate the difference among remotely sensed pixels. The wavelet-feature correlation ratio Markov [...] Read more.
The spectrums of one type of object under different conditions have the same features (up, down, protruding, concave) at the same spectral positions, which can be used as primary parameters to evaluate the difference among remotely sensed pixels. The wavelet-feature correlation ratio Markov clustering algorithm (WFCRMCA) for remotely sensed data is proposed based on an accurate description of abrupt spectral features and an optimized Markov clustering in the wavelet feather space. The peak points can be captured and identified by applying a wavelet transform to spectral data. The correlation ratio between two samples is a statistical calculation of the matched peak point positions on the wavelet feature within an adjustable spectrum domain or a range of wavelet scales. The evenly sampled data can be used to create class centers, depending on the correlation ratio threshold at each Markov step, accelerating the clustering speed by avoiding the computation of Euclidean distance for traditional clustering algorithms, such as K-means and ISODATA. Markov clustering applies several strategies, such as a simulated annealing method and gradually shrinking the clustering size, to control the clustering convergence. It can quickly obtain the best class centers at each clustering temperature. The experimental results of the Airborne Visible/Infrared Imaging Spectrometer (AVIRIS) and Thermal Mapping (TM) data have verified its acceptable clustering accuracy and high convergence velocity. Full article
(This article belongs to the Special Issue Novel Approaches for Remote Sensing Image Processing)
Show Figures

Figure 1

Figure 1
<p>Five points at different spatial positions within the same class have the same features at the exact spectral locations.</p>
Full article ">Figure 2
<p>The wavelet band-pass filter and four kinds of abrupt signals. (<b>a</b>–<b>d</b>) are the four critical signals: upward-maximal point, downward-minimal point, protruding crossing zero point, and concave crossing zero point. <span class="html-italic">ψ</span>(<span class="html-italic">t</span>) is the band-pass wavelet filter, (<b>a’</b>–<b>d’</b>) are the output of the four signals through the wavelet filter.</p>
Full article ">Figure 3
<p>(<b>a</b>) Closed set composed of five states. (<b>b</b>) Closed set with two states.</p>
Full article ">Figure 4
<p>Flow chart of WFCRMCA.</p>
Full article ">Figure 5
<p>(<b>a</b>) Mississippi TM 4th band image after gray balance. (<b>b</b>) Sook Lake AVIRIS 60th band image after gray balance.</p>
Full article ">Figure 6
<p>Mississippi TM image, WFCRMCA clustering results: (<b>a</b>–<b>i</b>) are the eight significant signals.</p>
Full article ">Figure 7
<p>Sook Lake AVIRIS image WFCRMCA clustering result. (<b>a</b>–<b>f</b>) are the seven significant signals.</p>
Full article ">
15 pages, 3534 KiB  
Communication
A Scalable Reduced-Complexity Compression of Hyperspectral Remote Sensing Images Using Deep Learning
by Sebastià Mijares i Verdú, Johannes Ballé, Valero Laparra, Joan Bartrina-Rapesta, Miguel Hernández-Cabronero and Joan Serra-Sagristà
Remote Sens. 2023, 15(18), 4422; https://doi.org/10.3390/rs15184422 - 8 Sep 2023
Cited by 2 | Viewed by 1533
Abstract
Two key hurdles to the adoption of Machine Learning (ML) techniques in hyperspectral data compression are computational complexity and scalability for large numbers of bands. These are due to the limited computing capacity available in remote sensing platforms and the high computational cost [...] Read more.
Two key hurdles to the adoption of Machine Learning (ML) techniques in hyperspectral data compression are computational complexity and scalability for large numbers of bands. These are due to the limited computing capacity available in remote sensing platforms and the high computational cost of compression algorithms for hyperspectral data, especially when the number of bands is large. To address these issues, a channel clusterisation strategy is proposed, which reduces the computational demands of learned compression methods for real scenarios and is scalable for different sources of data with varying numbers of bands. The proposed method is compatible with an embedded implementation for state-of-the-art on board hardware, a first for a ML hyperspectral data compression method. In terms of coding performance, our proposal surpasses established lossy methods such as JPEG 2000 preceded by a spectral Karhunen-Loève Transform (KLT), in clusters of 3 to 7 bands, achieving a PSNR improvement of, on average, 9 dB for AVIRIS and 3 dB for Hyperion images. Full article
(This article belongs to the Special Issue Recent Progress in Hyperspectral Remote Sensing Data Processing)
Show Figures

Figure 1

Figure 1
<p>Architecture from [<a href="#B6-remotesensing-15-04422" class="html-bibr">6</a>] with variable normalisation.</p>
Full article ">Figure 2
<p>PSNR results of models trained and tested for band-by-band compression of Landsat 8 OLI and AVIRIS images.</p>
Full article ">Figure 3
<p>Crop of AVIRIS scene f090710t01p00r09rdn-b band 110, compressed at 0.05 bits per sample (bps).</p>
Full article ">Figure 4
<p>PSNR performance of models trained and tested for compression in clusters of 3 bands of AVIRIS and Hyperion images.</p>
Full article ">Figure 5
<p>RSE performance of models trained and tested for compression in clusters of 3 bands of AVIRIS and Hyperion images.</p>
Full article ">Figure 6
<p>Spectral angle performance of models trained and tested for compression in clusters of 3 bands of AVIRIS and Hyperion images.</p>
Full article ">Figure 7
<p>Reconstruction of some selected bands from AVIRIS scene f090710t01p00r09rdn-b using our method, with zoom-in to better appreciate some od the differences. The overall image is compressed at 0.11 bps and overall reconstruction PSNR is 60.92 dB.</p>
Full article ">Figure 8
<p>Reconstruction of some selected bands from Hyperion scene EO1H1980182017066110K3 using our method, with zoom-in to better appreciate some od the differences. The overall image was compressed at 0.05 bps and overall reconstruction loss was 63.78 dB PSNR.</p>
Full article ">
16 pages, 10152 KiB  
Article
Effect of Spatial and Spectral Scaling on Joint Characterization of the Spectral Mixture Residual: Comparative Analysis of AVIRIS and WorldView-3 SWIR for Geologic Mapping in Anza-Borrego Desert State Park
by Jeffrey Price, Daniel Sousa and Francis J. Sousa
Sensors 2023, 23(15), 6742; https://doi.org/10.3390/s23156742 - 28 Jul 2023
Cited by 3 | Viewed by 1344
Abstract
A geologic map is both a visual depiction of the lithologies and structures occurring at the Earth’s surface and a representation of a conceptual model for the geologic history in a region. The work needed to capture such multifaced information in an accurate [...] Read more.
A geologic map is both a visual depiction of the lithologies and structures occurring at the Earth’s surface and a representation of a conceptual model for the geologic history in a region. The work needed to capture such multifaced information in an accurate geologic map is time consuming. Remote sensing can complement traditional primary field observations, geochemistry, chronometry, and subsurface geophysical data in providing useful information to assist with the geologic mapping process. Two novel sources of remote sensing data are particularly relevant for geologic mapping applications: decameter-resolution imaging spectroscopy (spectroscopic imaging) and meter-resolution multispectral shortwave infrared (SWIR) imaging. Decameter spectroscopic imagery can capture important mineral absorptions but is frequently unable to spatially resolve important geologic features. Meter-resolution multispectral SWIR images are better able to resolve fine spatial features but offer reduced spectral information. Such disparate but complementary datasets can be challenging to integrate into the geologic mapping process. Here, we conduct a comparative analysis of spatial and spectral scaling for two such datasets: one Airborne Visible/Infrared Imaging Spectrometer—Classic (AVIRIS-classic) flightline, and one WorldView-3 (WV3) scene, for a geologically complex landscape in Anza-Borrego Desert State Park, California. To do so, we use a two-stage framework that synthesizes recent advances in the spectral mixture residual and joint characterization. The mixture residual uses the wavelength-explicit misfit of a linear spectral mixture model to capture low variance spectral signals. Joint characterization utilizes nonlinear dimensionality reduction (manifold learning) to visualize spectral feature space topology and identify clusters of statistically similar spectra. For this study area, the spectral mixture residual clearly reveals greater spectral dimensionality in AVIRIS than WorldView (99% of variance in 39 versus 5 residual dimensions). Additionally, joint characterization shows more complex spectral feature space topology for AVIRIS than WorldView, revealing information useful to the geologic mapping process in the form of mineralogical variability both within and among mapped geologic units. These results illustrate the potential of recent and planned imaging spectroscopy missions to complement high-resolution multispectral imagery—along with field and lab observations—in planning, collecting, and interpreting the results from geologic field work. Full article
(This article belongs to the Section Remote Sensors)
Show Figures

Figure 1

Figure 1
<p>Index map. Anza-Borrego Desert State Park (white outline; both panels) spans regional geologic, topographic, and ecological gradients in California, USA. This study uses one AVIRIS flight line (orange) and one WorldView-3 image (blue). The analysis focused on the triple intersection (yellow rectangle; both panels) of the park boundary with AVIRIS flight line (orange) and WorldView-3 image (blue). True color basemap. The inset map shows the entire state of California; the green box overlay shows the extent of the left panel; and the gray box overlay shows the extent of the right panel.</p>
Full article ">Figure 2
<p>Study area (extent of yellow box in <a href="#sensors-23-06742-f001" class="html-fig">Figure 1</a>). True color AVIRIS image (<b>left</b>) with geologic unit boundary vectors overlaid (white; [<a href="#B25-sensors-23-06742" class="html-bibr">25</a>]), compared to SWIR false color composites of AVIRIS (<b>center</b>) and WorldView-3 (<b>right</b>). Compared to the true color imagery, geologic information is more apparent in the SWIR composites, particularly the carbonate absorptions in the marble unit (cyan; NE corner of all images). Note that AVIRIS is spatially coarser than WV3 (15 m versus 4 m) but spectrally finer (10 nm imaging spectroscopy vs. broadband multispectral). Linear 2% stretch applied to each image.</p>
Full article ">Figure 3
<p>Spectral feature space comparison for AVIRIS vs. WorldView-3 reflectance. PC 1 vs. 2 feature spaces of both sensors share similar topology and geometry. Endmember spectra (lower right for each sensor) correspond to the Palm Springs Formation (psf), marble (m), and quartz diorite (qd<sub>1</sub>) geologic units. The fourth endmember in WV3 corresponds to a different spatial subset of the pixels representing the quartz diorite geologic unit (qd<sub>2</sub>), likely related to intra-unit mineralogical variation within the unit that was not represented by the geologic map.</p>
Full article ">Figure 4
<p>Spectral mixture models and residuals. The results are shown for both AVIRIS (<b>A</b>,<b>C</b>) and WorldView (<b>B</b>,<b>D</b>). For spectral mixture models (<b>A</b>,<b>B</b>), red corresponds to the Palm Spring Formation (psf); blue corresponds to marble (m); and green corresponds to quartz diorite (qd<sub>1</sub>). For the spectral mixture residual (<b>C</b>,<b>D</b>), the spectral mixture model residual is shown at 2235, 1710, and 1550 nm (red, green, and blue, respectively) wavelengths. Potentially useful geologic information, indicated by the diversity of colors, is clearly present in the residual both within and among geologic units. A linear 2% stretch was applied to each image.</p>
Full article ">Figure 5
<p>Mixture residual spectral feature space. Regions associated with different geologic features are identifiable for both sensors, but substantially more spectral information can be obtained from AVIRIS than WV3. Specifically, AVIRIS captures differences in VNIR curvature and amplitude, and narrow SWIR absorptions, which are not resolved by WV3. The reflectance spectra for endmembers identified from the MR feature space (lower right for each sensor) correspond to W, X, Y, and Z in the AVIRIS image and A, B, and C in the WorldView-3 image.</p>
Full article ">Figure 6
<p>Partition of variance. Normalized eigenvalue distributions are shown for surface reflectance (<b>left</b>) versus mixture residual (<b>right</b>), as well as for AVIRIS (blue) and WorldView-3 (orange). Inset shows the cumulative variance. AVIRIS reflectance shows greater spectral dimensionality than WorldView-3, with 99% of image variance requiring 7 dimensions for AVIRIS versus only 2 dimensions for WV3. Between-sensor differences in dimensionality are further pronounced in the mixture residual, with AVIRIS capturing 90% variance in 7 bands and 99% variance in 39 bands, but with WorldView-3 capturing 90% variance in 4 bands and 99% variance in 5 bands. WorldView-3 residual eigenvalues are numerically indistinct from zero after the 5th dimension.</p>
Full article ">Figure 7
<p>UMAP feature space comparison. For both AVIRIS and WorldView-3 reflectance imagery, nonlinear dimensionality reduction via UMAP complements linear decomposition by PCA, revealing considerably more apparent clusters in regions shown to be diffuse in low-order PC space. EMs follow the same color scheme as previous figures (red: PSF, cyan: qd<sub>1</sub>, green: m, and yellow: qd<sub>2</sub>). Note the enhanced separability of red and yellow clusters in AVIRIS relative to WV3.</p>
Full article ">Figure 8
<p>Joint characterization of AVIRIS reflectance. UMAP 1 vs. 2 feature space (<b>top right</b>). The clusters were manually selected on the UMAP space and overlaid on PCA space (<b>middle right</b>) and joint space (<b>bottom right</b>). The clusters identified from UMAP also generally cluster in PC space but are much more diffuse. These clusters also persist in joint space, with the additional benefit of physical interpretability provided by the abundance of one of the spectral mixture model endmembers. The true color AVIRIS image (<b>left</b>) with UMAP clusters and geologic unit boundaries overlaid (white) shows the relationship between statistical clustering in UMAP feature space and geographic clustering of geologic features. The EMs are labeled on each scatter plot.</p>
Full article ">Figure 9
<p>Intra-unit heterogeneity. (<b>A</b>) Geologic map units overlaid on the study area (<b>left</b>). Granite (Gr) unit highlighted in yellow on the AVIRIS true color image (<b>center left</b>). The enlarged view of the Gr unit on AVIRIS (<b>center right</b>) and WorldView-3 (<b>right</b>) highlights differences in spatial resolution. Single-pixel spectra are marked by red circles. (<b>B</b>) The average spectrum of the Gr unit (white line) is compared to the 12 single-pixel spectra selected within the unit for the reflectance (<b>top row</b>) and residual (<b>bottom row</b>) of both AVIRIS and WorldView-3. The left column row displays the full spectral extent of AVIRIS, while the middle column trims the AVIRIS spectra to the extent of the WorldView-3 column on the right. Differences between the single-pixel reflectance spectra and the Gr unit average spectrum are more clearly identifiable in the residual spectra than in the reflectance spectra and in the AVIRIS spectra than in the WorldView-3 spectra. All reflectance images are shown in true color.</p>
Full article ">
12 pages, 1574 KiB  
Article
Hyperspectral Bare Soil Index (HBSI): Mapping Soil Using an Ensemble of Spectral Indices in Machine Learning Environment
by Eric Ariel L. Salas and Sakthi Subburayalu Kumaran
Land 2023, 12(7), 1375; https://doi.org/10.3390/land12071375 - 10 Jul 2023
Cited by 2 | Viewed by 2985
Abstract
Spectral remote-sensing indices based on visible, NIR, and SWIR wavelengths are useful in predicting spatial patterns of bare soil. However, identifying an effective combination of informative wavelengths or spectral indices for mapping bare soil in a complex urban/agricultural region is still a challenge. [...] Read more.
Spectral remote-sensing indices based on visible, NIR, and SWIR wavelengths are useful in predicting spatial patterns of bare soil. However, identifying an effective combination of informative wavelengths or spectral indices for mapping bare soil in a complex urban/agricultural region is still a challenge. In this study, we developed a new bare-soil index, the Hyperspectral Bare Soil Index (HBSI), to improve the accuracy of bare-soil remote-sensing mapping. We tested the HBSI using the high-spectral-resolution AVIRIS-NG and Sentinel-2 multispectral images. We applied an ensemble modeling approach, consisting of random forest (RF) and support vector machine (SVM), to classify bare soil. We found that the HBSI outperformed other existing bare-soil indices with over 91% accuracy for Sentinel-2 and AVIRIS-NG. Furthermore, the combination of the HBSI and the normalized difference vegetation index (NDVI) showed a better performance in bare-soil classification, with >92% accuracy for Sentinel-2 and >97% accuracy for AVIRIS-NG images. Also, the RF-SVM ensemble surpassed the performance of the individual models. The novelty of HBSI is due to its development, since it utilizes the blue band in addition to the NIR and SWIR2 bands from the high-spectral-resolution data from AVIRIS-NG to improve the accuracy of bare-soil mapping. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Map showing the location of the study site in the Anand District in Gujarat, India. Examples of farmlands with sparse vegetation and bare-soil conditions are shown in (<b>b</b>,<b>c</b>), respectively.</p>
Full article ">Figure 2
<p>Bare soil-urban map derived using an ensemble of RF and SVM from: (<b>a</b>) AVIRIS-NG; (<b>b</b>) Sentinel-2 images; and (<b>c</b>) magnified section to highlight the separation of classes.</p>
Full article ">
19 pages, 42392 KiB  
Article
Detection of Solar Photovoltaic Power Plants Using Satellite and Airborne Hyperspectral Imaging
by Christoph Jörges, Hedwig Sophie Vidal, Tobias Hank and Heike Bach
Remote Sens. 2023, 15(13), 3403; https://doi.org/10.3390/rs15133403 - 5 Jul 2023
Cited by 8 | Viewed by 4573
Abstract
Solar photovoltaic panels (PV) provide great potential to reduce greenhouse gas emissions as a renewable energy technology. The number of solar PV has increased significantly in recent years and is expected to increase even further. Therefore, accurate and global mapping and monitoring of [...] Read more.
Solar photovoltaic panels (PV) provide great potential to reduce greenhouse gas emissions as a renewable energy technology. The number of solar PV has increased significantly in recent years and is expected to increase even further. Therefore, accurate and global mapping and monitoring of PV modules with remote sensing methods is important for predicting energy production potentials, revealing socio-economic drivers, supporting urban planning, and estimating ecological impacts. Hyperspectral imagery provides crucial information to identify PV modules based on their physical absorption and reflection properties. This study investigated spectral signatures of spaceborne PRISMA data of 30 m low resolution for the first time, as well as airborne AVIRIS-NG data of 5.3 m medium resolution for the detection of solar PV. The study region is located around Irlbach in southern Germany. A physics-based approach using the spectral indices nHI, NSPI, aVNIR, PEP, and VPEP was used for the classification of the hyperspectral images. By validation with a solar PV ground truth dataset of the study area, a user’s accuracy of 70.53% and a producer’s accuracy of 88.06% for the PRISMA hyperspectral data, and a user’s accuracy of 65.94% and a producer’s accuracy of 82.77% for AVIRIS-NG were achieved. Full article
Show Figures

Figure 1

Figure 1
<p>Workflow of the methodological approach of this study.</p>
Full article ">Figure 2
<p>Map of the study area showing (<b>A</b>) the airborne HSI data of AVIRIS-NG acquired as part of ESA CHIME mission preparations on 30 May 2021, including the flight lines. The HSI data are shown in true color. Ground truth data of solar PV modules with more than 10 kWp are shown on the map (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS). (<b>B</b>) provides an overview of the study area in Bavaria, south-east Germany.</p>
Full article ">Figure 3
<p>Spectral signature of polycrystalline solar PV modules from hyperspectral AVIRIS-NG (orange line) and multispectral Sentinel-2 (blue dots) data. While the AVIRIS-NG covers the spectrum from 377 to 2500 nm with a spectral resolution of 5 nm, the nine bands of Sentinel-2 range from 492 to 2202 nm.</p>
Full article ">Figure 4
<p>Overview of the five spectral indices nHI, NSPI, aVNIR, PEP, and VPEP that have been calculated to see their applicability to detect PV modules in airborne and spaceborne data with a spatial resolution of 5.3 m (AVIRIS) and 30 m (PRISMA) per pixel. The presented PV spectrum origins from PV modules installed on top of an industrial building and was acquired with AVIRIS-NG.</p>
Full article ">Figure 5
<p>Classification results for the airborne HSI data of AVIRIS-NG acquired as part of ESA CHIME mission preparations on 30 May 2021. The HSI data are shown in true color for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>). Ground truth data of solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown on the map.</p>
Full article ">Figure 6
<p>nHI calculated from the AVIRIS-NG data acquired on 30 May 2021 for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>) with the solar PV park Gänsdorf in detail map (<b>C</b>). Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown.</p>
Full article ">Figure 7
<p>NSPI calculated from the AVIRIS-NG data acquired on 30 May 2021 for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>) with the solar PV park Gänsdorf in detail map (<b>C</b>). Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown.</p>
Full article ">Figure 8
<p>aVNIR, PEP, and VPEP calculated from the AVIRIS-NG data acquired on 30 May 2021. Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown.</p>
Full article ">Figure 9
<p>Classification results for the spaceborne HSI data of PRISMA acquired from ASI on 26 March 2022 for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>) with the solar PV park Gänsdorf in detail map (<b>C</b>). The HSI data are shown in true color. Ground truth data of solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown on the map. PRISMA Product-© Italian Space Agency (ASI) 2022. All rights reserved.</p>
Full article ">Figure 10
<p>nHI calculated from the PRISMA data acquired on 26 March 2022 for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>) with the solar PV park Gänsdorf in detail map (<b>C</b>). Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown. PRISMA Product-© Italian Space Agency (ASI) 2022. All rights reserved.</p>
Full article ">Figure 11
<p>NSPI calculated from the PRISMA data acquired on 26 March 2022 for the entire study area (<b>A</b>), as well as for three details of the map (<b>B</b>–<b>D</b>) with the solar PV park Gänsdorf in detail map (<b>C</b>). Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown. PRISMA Product-© Italian Space Agency (ASI) 2022. All rights reserved.</p>
Full article ">Figure 12
<p>aVNIR, PEP, and VPEP calculated from the PRISMA data acquired on 26 March 2022. Solar PV modules with more than 10 kWp (Source: Bavarian Agency for the Environment, Energie-Atlas Bayern: Photovoltaikanlagen-WMS) are shown. PRISMA Product-© Italian Space Agency (ASI) 2022. All rights reserved.</p>
Full article ">Figure 13
<p>Comparison of spectral signatures of vegetation on a field, a forest, and a PV module with surrounding vegetation. The rapid increase in the red edge is also abundant in the mixed PV spectrum, thus missing the typically low reflectance of PV modules up to 1000 nm. The S-curve between 763 and 873 nm is obtained only in the mixed PV spectrum.</p>
Full article ">
13 pages, 5203 KiB  
Article
Anomaly Detection of Remote Sensing Images Based on the Channel Attention Mechanism and LRX
by Huinan Guo, Hua Wang, Xiaodong Song and Zhongling Ruan
Appl. Sci. 2023, 13(12), 6988; https://doi.org/10.3390/app13126988 - 9 Jun 2023
Cited by 2 | Viewed by 1891
Abstract
Anomaly detection of remote sensing images has gained significant attention in remote sensing image processing due to their rich spectral information. The Local RX (LRX) algorithm, derived from the Reed–Xiaoli (RX) algorithm, is a hyperspectral anomaly detection method that focuses on identifying anomalous [...] Read more.
Anomaly detection of remote sensing images has gained significant attention in remote sensing image processing due to their rich spectral information. The Local RX (LRX) algorithm, derived from the Reed–Xiaoli (RX) algorithm, is a hyperspectral anomaly detection method that focuses on identifying anomalous pixels in hyperspectral images by exploiting local statistics and background modeling. However, it is still susceptible to the noises in the Hyperspectral Images (HSIs), which limits its detection performance. To address this problem, a hyperspectral anomaly detection algorithm based on channel attention mechanism and LRX is proposed in this paper. The HSI is feed into the auto-encoder network that is constrained by the channel attention module to generate a more representative reconstructed image that better captures the characteristics of different land covers and has less noises. The channel attention module in the auto-encoder network aims to explore the effective spectral bands corresponding to different land covers. Subsequently, the LRX algorithm is utilized for anomaly detection on the reconstructed image obtained from the auto-encoder network with the channel attention mechanism, which avoids the influence of noises on the anomaly detection results and improves the anomaly detection performance. The experiments are conducted on three HSIs to verify the performance of the proposed method. The proposed hyperspectral anomaly detection method achieves higher Area Under Curve (AUC) values of 0.9871, 0.9916 and 0.9642 on HYDICE urban dataset, AVIRIS aircraft dataset and Salinas Valley dataset, respectively, compared with other six methods. The experimental results demonstrate that the proposed algorithm has better anomaly detection performance than LRX and other algorithms. Full article
(This article belongs to the Special Issue AI-Based Image Processing)
Show Figures

Figure 1

Figure 1
<p>Network structure of the proposed hyperspectral anomaly detection algorithm based on the channel attention mechanism and LRX.</p>
Full article ">Figure 2
<p>(<b>a</b>) The pseudo-color image of HYDICE urban dataset. (<b>b</b>) The ground truth.</p>
Full article ">Figure 3
<p>(<b>a</b>) The pseudo-color image of AVIRIS airplane dataset. (<b>b</b>) The ground truth.</p>
Full article ">Figure 4
<p>(<b>a</b>) The pseudo-color image of Salinas Valley dataset. (<b>b</b>) The ground truth.</p>
Full article ">Figure 5
<p>ROC curves of six comparative algorithms and the proposed algorithm on three data sets: (<b>a</b>) HYDICE urban data; (<b>b</b>) AVIRIS airplane data; (<b>c</b>) Salinas scene.</p>
Full article ">Figure 6
<p>Visualized anomaly detection results of different algorithms on the HYDICE urban dataset: (<b>a</b>) LRX; (<b>b</b>) BJSR; (<b>c</b>) LRaSMD; (<b>d</b>) MCRD; (<b>e</b>) AE-based; (<b>f</b>) AUTO-AD; (<b>g</b>) ours; (<b>h</b>) ground truth.</p>
Full article ">Figure 7
<p>Visualized anomaly detection results of different algorithms on AVIRIS airplane data set: (<b>a</b>) LRX; (<b>b</b>) BJSR; (<b>c</b>) LRaSMD; (<b>d</b>) MCRD; (<b>e</b>) AE-based; (<b>f</b>) AUTO-AD; (<b>g</b>) ours; (<b>h</b>) ground truth.</p>
Full article ">Figure 8
<p>Visualized anomaly detection results of different algorithms on Salinas Valley data set: (<b>a</b>) LRX; (<b>b</b>) BJSR; (<b>c</b>) LRaSMD; (<b>d</b>) MCRD; (<b>e</b>) AE-based; (<b>f</b>) AUTO-AD; (<b>g</b>) ours; (<b>h</b>) ground truth.</p>
Full article ">
20 pages, 10206 KiB  
Article
Region Expansion of a Hyperspectral-Based Mineral Map Using Random Forest Classification with Multispectral Data
by Hideki Tsubomatsu and Hideyuki Tonooka
Minerals 2023, 13(6), 754; https://doi.org/10.3390/min13060754 - 31 May 2023
Cited by 1 | Viewed by 1697
Abstract
Observation images from hyperspectral (HS) sensors on satellites and aircraft can be used to map minerals in greater detail than those from multispectral (MS) sensors. However, the coverage of HS images is much less than that of MS images, so there are often [...] Read more.
Observation images from hyperspectral (HS) sensors on satellites and aircraft can be used to map minerals in greater detail than those from multispectral (MS) sensors. However, the coverage of HS images is much less than that of MS images, so there are often cases where MS images cover the entire area of interest while HS images cover only a part of it. In this study, we propose a new method to more reasonably expand the mineral map of an HS image with an MS image in such cases. The method uses various mineral indices from the MS image and MS sensor’s band values as the input and HS image-based mineral classes as the output. Random forest (RF) two-class classification is then applied iteratively to determine the distribution of each mineral in turn, starting with the minerals that are most consistent with the HS image-based mineral map. The method also involves the correction of misalignment between HS and MS images and the selection of input variables by RF multiclass classification. The method was evaluated in comparison with other methods in the Cuprite area, Nevada, using the Airborne Visible/Infrared Imaging Spectrometer (AVIRIS) and Hyperspectral Imager Suite (HISUI) as HS sensors and the Advanced Spaceborne Thermal Emission and Reflection Radiometer (ASTER) as MS sensors. As a result, all of the evaluated region-expansion methods with an HS–MS image pair, including the proposed method, showed better performance than the method using only an MS image. The proposed method had the highest performance, and the inter-mineral averages of the F1-scores for the overlap and non-overlap areas were 85.98% and 46.46% for the AVIRIS–ASTER image pair and 82.78% and 42.60% for the HISUI–ASTER image pair, respectively. Although the performance in the non-overlap region was lower than in the overlap region, the method showed high precision and high accuracy for almost all minerals, including minerals with only a few pixels. Misalignment between the HS–MS images is a factor that degrades accuracy and requires precise alignment, but the misalignment correction in the proposed method could suppress the effect of misalignment. Validation studies using different regions and different sensors will be carried out in the future. Full article
Show Figures

Figure 1

Figure 1
<p>Processing flow of the proposed method.</p>
Full article ">Figure 2
<p>Cuprite area, Nevada, USA, used as the study area.</p>
Full article ">Figure 3
<p>Comparison of mineral maps obtained from a single AVIRIS image and mineral maps from each of the extension methods using ASTER images: (<b>a</b>) AVIRIS-based map and the extended maps by (<b>b</b>) proposed method, (<b>c</b>) method A, (<b>d</b>) method B, (<b>e</b>) improved HT method, and (<b>f</b>) MS-based method. The white dotted box in (<b>a</b>) indicates the overlap region.</p>
Full article ">Figure 4
<p>Comparison of mineral maps obtained from a single HISUI image and mineral maps from each of the extension methods using ASTER images: (<b>a</b>) HISUI-based map and the extended maps by (<b>b</b>) proposed method, (<b>c</b>) method A, (<b>d</b>) method B, (<b>e</b>) improved HT method, and (<b>f</b>) MS-based method. The white dotted box in (<b>a</b>) indicates the overlap region.</p>
Full article ">Figure 5
<p>Mineral maps for the case of misalignment between AVIRIS and ASTER images: (<b>a</b>) AVIRIS-based map and the extended maps by (<b>b</b>) proposed method, (<b>c</b>) method A, (<b>d</b>) method B, and (<b>e</b>) improved HT method. The white dotted box in (<b>a</b>) indicates the overlap region.</p>
Full article ">
Back to TopTop