[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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (499)

Search Parameters:
Keywords = mixed pixels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
31 pages, 19893 KiB  
Article
A Low-Measurement-Cost-Based Multi-Strategy Hyperspectral Image Classification Scheme
by Yu Bai, Dongmin Liu, Lili Zhang and Haoqi Wu
Sensors 2024, 24(20), 6647; https://doi.org/10.3390/s24206647 - 15 Oct 2024
Viewed by 288
Abstract
The cost of hyperspectral image (HSI) classification primarily stems from the annotation of image pixels. In real-world classification scenarios, the measurement and annotation process is both time-consuming and labor-intensive. Therefore, reducing the number of labeled pixels while maintaining classification accuracy is a key [...] Read more.
The cost of hyperspectral image (HSI) classification primarily stems from the annotation of image pixels. In real-world classification scenarios, the measurement and annotation process is both time-consuming and labor-intensive. Therefore, reducing the number of labeled pixels while maintaining classification accuracy is a key research focus in HSI classification. This paper introduces a multi-strategy triple network classifier (MSTNC) to address the issue of limited labeled data in HSI classification by improving learning strategies. First, we use the contrast learning strategy to design a lightweight triple network classifier (TNC) with low sample dependence. Due to the construction of triple sample pairs, the number of labeled samples can be increased, which is beneficial for extracting intra-class and inter-class features of pixels. Second, an active learning strategy is used to label the most valuable pixels, improving the quality of the labeled data. To address the difficulty of sampling effectively under extremely limited labeling budgets, we propose a new feature-mixed active learning (FMAL) method to query valuable samples. Fine-tuning is then used to help the MSTNC learn a more comprehensive feature distribution, reducing the model’s dependence on accuracy when querying samples. Therefore, the sample quality is improved. Finally, we propose an innovative dual-threshold pseudo-active learning (DSPAL) strategy, filtering out pseudo-label samples with both high confidence and uncertainty. Extending the training set without increasing the labeling cost further improves the classification accuracy of the model. Extensive experiments are conducted on three benchmark HSI datasets. Across various labeling ratios, the MSTNC outperforms several state-of-the-art methods. In particular, under extreme small-sample conditions (five samples per class), the overall accuracy reaches 82.97% (IP), 87.94% (PU), and 86.57% (WHU). Full article
Show Figures

Figure 1

Figure 1
<p>The overall architecture of the proposed MSTN for HIS classification. <span class="html-fig-inline" id="sensors-24-06647-i001"><img alt="Sensors 24 06647 i001" src="/sensors/sensors-24-06647/article_deploy/html/images/sensors-24-06647-i001.png"/></span>: encoding layer vector of the sample; <span class="html-fig-inline" id="sensors-24-06647-i002"><img alt="Sensors 24 06647 i002" src="/sensors/sensors-24-06647/article_deploy/html/images/sensors-24-06647-i002.png"/></span>: the classifier output vector of the sample; ⊕: add selected sample to the train-set; ⊖: remove selected sample from the test-set; filter 1: a filter used in active learning strategies to screen samples using the FMAX method; filter 2: a filter used in the pseudo-active learning strategy, which includes two filters that filter high-confidence samples and high-uncertainty samples, respectively.</p>
Full article ">Figure 2
<p>TNC detailed model structure. <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> </mrow> </semantics></math>@<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>1</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>2</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> </mrow> </semantics></math>: <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>0</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> </mrow> </semantics></math> denote the size of the convolution kernel; <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>1</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>2</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> <msubsup> <mrow> <mi>k</mi> </mrow> <mrow> <mn>3</mn> </mrow> <mrow> <mi>i</mi> </mrow> </msubsup> </mrow> </semantics></math> represents the three dimensions of the input data; <math display="inline"><semantics> <mrow> <mi>i</mi> </mrow> </semantics></math> represents the <span class="html-italic">i</span>-th layer convolution.</p>
Full article ">Figure 3
<p>Feature mixing-based active learning (FMAL) illustration. ⊕: add selected sample to the train-set; ⊖: remove selected sample from the test-set; Repeat: number of iterations; <span class="html-fig-inline" id="sensors-24-06647-i003"><img alt="Sensors 24 06647 i003" src="/sensors/sensors-24-06647/article_deploy/html/images/sensors-24-06647-i003.png"/></span>: training TNC with train-set.</p>
Full article ">Figure 4
<p>The proposed DSPAL architecture. A filter is established to screen samples classified by the classifier. The samples are labeled using the predicted values of the classifier, and the qualified pseudo-labeled samples are then added to <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>D</mi> </mrow> <mrow> <mi>t</mi> <mi>r</mi> <mi>a</mi> <mi>i</mi> <mi>n</mi> </mrow> </msup> </mrow> </semantics></math> for iterative training.</p>
Full article ">Figure 5
<p>Indian Pines dataset: (<b>a</b>) false-color map; (<b>b</b>) ground-truth map. The numbers in parentheses represent the total number of samples in each class.</p>
Full article ">Figure 6
<p>Pavia University dataset: (<b>a</b>) false-color map; (<b>b</b>) ground-truth map. The numbers in parentheses represent the total number of samples in each class.</p>
Full article ">Figure 7
<p>Wuhan Han Chuan dataset: (<b>a</b>) false-color map; (<b>b</b>) ground-truth map. The numbers in parentheses represent the total number of samples in each class.</p>
Full article ">Figure 8
<p>Classification maps of the different models for the Indian Pines dataset: (<b>a</b>) 3DCNN; (<b>b</b>) DFSL-NN; (<b>c</b>) DFSL-SVM; (<b>d</b>) Gia-CFSL; (<b>e</b>) DBDAFSL; (<b>f</b>) CapsGLOM; (<b>g</b>) MSTNC.</p>
Full article ">Figure 9
<p>Classification maps of the different models for the Pavia University dataset: (<b>a</b>) 3DCNN; (<b>b</b>) DFSL-NN; (<b>c</b>) DFSL-SVM; (<b>d</b>) Gia-CFSL; (<b>e</b>) DBDAFSL; (<b>f</b>) CapsGLOM; (<b>g</b>) MSTNC.</p>
Full article ">Figure 10
<p>Classification maps of the different models for the WHU-Hi-HanChuan dataset: (<b>a</b>) 3DCNN; (<b>b</b>) DFSL-NN; (<b>c</b>) DFSL-SVM; (<b>d</b>) Gia-CFSL; (<b>e</b>) DBDAFSL; (<b>f</b>) CapsGLOM; (<b>g</b>) MSTNC.</p>
Full article ">Figure 11
<p>Evolution of OA as a function of the number of training samples per class: (<b>a</b>) IP dataset; (<b>b</b>) PU dataset; (<b>c</b>) WHU dataset.</p>
Full article ">Figure 12
<p>Evolution of AA as a function of the number of training samples per class: (<b>a</b>) IP dataset; (<b>b</b>) PU dataset; (<b>c</b>) WHU dataset.</p>
Full article ">Figure 13
<p>Evolution of Kappa as a function of the number of training samples per class: (<b>a</b>) IP dataset; (<b>b</b>) PU dataset; (<b>c</b>) WHU dataset.</p>
Full article ">Figure 14
<p>Training time and testing time of different methods.</p>
Full article ">Figure 15
<p>Ablation experiments of the triplet contrastive learning.</p>
Full article ">Figure 16
<p>Ablation experiments of the proposed method on.</p>
Full article ">Figure 17
<p>Comparison of accuracy of three different sample selection methods: (<b>a</b>) OA; (<b>b</b>) AA; (<b>c</b>) Kappa.</p>
Full article ">Figure 18
<p>Classification maps of different methods for Indian Pines. Seven labeled samples were selected for each class. Light blue dots are randomly selected samples. The red dots is samples selected with a specific method. Dark blue dots are samples that were incorrectly predicted. (<b>a</b>) Random method; (<b>b</b>) BvSB method; (<b>c</b>) FMAL method.</p>
Full article ">Figure 19
<p>Analysis of the role of iteration in active learning methods. The data in parentheses in the horizontal axis represent the number of iterations of ALn. (<b>a</b>) OA; (<b>b</b>) AA; (<b>c</b>) Kappa.</p>
Full article ">Figure 20
<p>Effect of the number of DSPAL iterations on accuracy: (<b>a</b>) IP dataset; (<b>b</b>) PU dataset; (<b>c</b>) WHU dataset.</p>
Full article ">
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 292
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 ">
17 pages, 9539 KiB  
Article
A Chaos-Based Encryption Algorithm to Protect the Security of Digital Artwork Images
by Li Shi, Xiangjun Li, Bingxue Jin and Yingjie Li
Mathematics 2024, 12(20), 3162; https://doi.org/10.3390/math12203162 - 10 Oct 2024
Viewed by 276
Abstract
Due to the security weaknesses of chaos-based pseudorandom number generators, in this paper, a new pseudorandom number generator (PRNG) based on mixing three-dimensional variables of a cat chaotic map is proposed. A uniformly distributed chaotic sequence by a logistic map is used in [...] Read more.
Due to the security weaknesses of chaos-based pseudorandom number generators, in this paper, a new pseudorandom number generator (PRNG) based on mixing three-dimensional variables of a cat chaotic map is proposed. A uniformly distributed chaotic sequence by a logistic map is used in the mixing step. Both statistical tests and a security analysis indicate that our PRNG has good randomness and is more complex than any one-dimensional variable of a cat map. Furthermore, a new image encryption algorithm based on the chaotic PRNG is provided to protect the content of artwork images. The core of the algorithm is to use the sequence generated by the pseudorandom number generator to achieve the process of disruption and diffusion of the image pixels, so as to achieve the effect of obfuscation and encryption of the image content. Several security tests demonstrate that this image encryption algorithm has a high security level. Full article
(This article belongs to the Special Issue Chaos-Based Secure Communication and Cryptography, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The state variables of cat chaotic map. (<b>a</b>) <span class="html-italic">xy</span>-dimensional; (<b>b</b>) <span class="html-italic">xz</span>-dimensional; (<b>c</b>) <span class="html-italic">yz</span>-dimensional; (<b>d</b>) chaotic attractor.</p>
Full article ">Figure 2
<p>The main frame of our PRNG.</p>
Full article ">Figure 3
<p>Comparison of ApEn for sequences {<span class="html-italic">b<sub>i</sub></span>}, {<span class="html-italic">x<sub>i</sub></span>}, {<span class="html-italic">y<sub>i</sub></span>}, and {<span class="html-italic">z<sub>i</sub></span>}.</p>
Full article ">Figure 4
<p>Linear complexity.</p>
Full article ">Figure 5
<p>Encryption and decryption tests: (<b>a</b>,<b>d</b>,<b>g</b>) denotes the original image; (<b>b</b>,<b>e</b>,<b>h</b>) denotes the encrypted image; (<b>c</b>,<b>f</b>,<b>i</b>) denotes the decrypted image.</p>
Full article ">Figure 6
<p>Histograms: (<b>a</b>–<b>c</b>) original images; (<b>d</b>–<b>f</b>) cryptographic images (the red, green and blue appearing in the diagram correspond to the three RGB colour channels).</p>
Full article ">Figure 7
<p>Pixel correlation analysis: (<b>a</b>–<b>i</b>) denote the pixel correlations of the original images of images 1–3 in different colour channels, horizontally, vertically and diagonally; (<b>j</b>–<b>r</b>) correspond to the pixel correlations of their secret images in different colour channels, horizontally, vertically and diagonally.</p>
Full article ">Figure 8
<p>Key sensitive analysis: (<b>a</b>) correctly decrypted image; (<b>b</b>) x<sub>0</sub> + 2<sup>−12</sup>; (<b>c</b>) y<sub>0</sub> + 2<sup>−12</sup>; (<b>d</b>) z<sub>0</sub> + 2<sup>−12</sup>.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>c</b>) denote the corresponding decrypted images after adding 5% salt and pepper noise to the three encrypted images, respectively.</p>
Full article ">
19 pages, 11653 KiB  
Article
Influence of Vegetation Phenology on the Temporal Effect of Crop Fractional Vegetation Cover Derived from Moderate-Resolution Imaging Spectroradiometer Nadir Bidirectional Reflectance Distribution Function–Adjusted Reflectance
by Yinghao Lin, Tingshun Fan, Dong Wang, Kun Cai, Yang Liu, Yuye Wang, Tao Yu and Nianxu Xu
Agriculture 2024, 14(10), 1759; https://doi.org/10.3390/agriculture14101759 - 5 Oct 2024
Viewed by 393
Abstract
Moderate-Resolution Imaging Spectroradiometer (MODIS) Nadir Bidirectional Reflectance Distribution Function (BRDF)-Adjusted Reflectance (NBAR) products are being increasingly used for the quantitative remote sensing of vegetation. However, the assumption underlying the MODIS NBAR product’s inversion model—that surface anisotropy remains unchanged over the 16-day retrieval period—may [...] Read more.
Moderate-Resolution Imaging Spectroradiometer (MODIS) Nadir Bidirectional Reflectance Distribution Function (BRDF)-Adjusted Reflectance (NBAR) products are being increasingly used for the quantitative remote sensing of vegetation. However, the assumption underlying the MODIS NBAR product’s inversion model—that surface anisotropy remains unchanged over the 16-day retrieval period—may be unreliable, especially since the canopy structure of vegetation undergoes stark changes at the start of season (SOS) and the end of season (EOS). Therefore, to investigate the MODIS NBAR product’s temporal effect on the quantitative remote sensing of crops at different stages of the growing seasons, this study selected typical phenological parameters, namely SOS, EOS, and the intervening stable growth of season (SGOS). The PROBA-V bioGEOphysical product Version 3 (GEOV3) Fractional Vegetation Cover (FVC) served as verification data, and the Pearson correlation coefficient (PCC) was used to compare and analyze the retrieval accuracy of FVC derived from the MODIS NBAR product and MODIS Surface Reflectance product. The Anisotropic Flat Index (AFX) was further employed to explore the influence of vegetation type and mixed pixel distribution characteristics on the BRDF shape under different stages of the growing seasons and different FVC; that was then combined with an NDVI spatial distribution map to assess the feasibility of using the reflectance of other characteristic directions besides NBAR for FVC correction. The results revealed the following: (1) Generally, at the SOSs and EOSs, the differences in PCCs before vs. after the NBAR correction mainly ranged from 0 to 0.1. This implies that the accuracy of FVC derived from MODIS NBAR is lower than that derived from MODIS Surface Reflectance. Conversely, during the SGOSs, the differences in PCCs before vs. after the NBAR correction ranged between –0.2 and 0, suggesting the accuracy of FVC derived from MODIS NBAR surpasses that derived from MODIS Surface Reflectance. (2) As vegetation phenology shifts, the ensuing differences in NDVI patterning and AFX can offer auxiliary information for enhanced vegetation classification and interpretation of mixed pixel distribution characteristics, which, when combined with NDVI at characteristic directional reflectance, could enable the accurate retrieval of FVC. Our results provide data support for the BRDF correction timescale effect of various stages of the growing seasons, highlighting the potential importance of considering how they differentially influence the temporal effect of NBAR corrections prior to monitoring vegetation when using the MODIS NBAR product. Full article
Show Figures

Figure 1

Figure 1
<p>Spatial extent of the Wancheng District study area (in Henan Province, China). (<b>a</b>) Map of land cover types showing the location of sampling points across the study area. This map came from MCD12Q1 (v061). (<b>b</b>–<b>d</b>) True-color images of the three mixed pixels, obtained from Sentinel-2. The distribution characteristics are as follows: crops above with buildings below (<b>b</b>); crops below with buildings above (<b>c</b>); and buildings in the upper-left corner, crops in the remainder (<b>d</b>).</p>
Full article ">Figure 2
<p>Monthly average temperature and monthly total precipitation in the study area, from 2017 to 2021.</p>
Full article ">Figure 3
<p>Data processing flow chart. The green rectangles from top to the bottom represent three steps: crop phenological parameters extraction with TIMESAT; Fractional Vegetation Cover (FVC) derived from MOD09GA and MCD43A4; and accuracy evaluation, respectively. Blue solid rectangles refer to a used product or derived results, while blue dashed rectangles refer to the software or model used in this study. NDVI<sub>MOD09GA</sub>: NDVI derived from MOD09GA, NDVI<sub>MCD43A4</sub>: NDVI derived from MCD43A4, FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA, FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. PCC<sub>MOD09GA</sub>: Pearson correlation coefficient (PCC) calculated for FVC<sub>MOD09GA</sub> and GEOV3 FVC, PCC<sub>MCD43A4</sub>: PCC calculated for FVC<sub>MCD43A4</sub> and GEOV3 FVC.</p>
Full article ">Figure 4
<p>NDVI and EVI time series fitted curves and phenological parameters of crops. SOS: start of season; EOS: end of season; SGOS: stable growth of season.</p>
Full article ">Figure 5
<p>Spatial distribution of Fractional Vegetation Cover (FVC) derived from MOD09GA and MCD43A4, and the difference images of FVC. FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA, FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. (<b>a</b>–<b>c</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 15 November 2020, respectively; (<b>d</b>–<b>f</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 10 February 2021, respectively; (<b>g</b>–<b>i</b>) FVC derived from MOD09GA, MCD43A4, and the difference between FVC<sub>MOD09GA</sub> and FVC<sub>MCD43A4</sub> on 30 September 2021, respectively.</p>
Full article ">Figure 6
<p>Pearson correlation coefficients (PCCs) of Fractional Vegetation Cover (FVC) derived before and after the NBAR correction with GEOV3 FVC at different stages of the growing seasons. FVC<sub>MOD09GA</sub>: FVC derived from MOD09GA. FVC<sub>MCD43A4</sub>: FVC derived from MCD43A4. PCC<sub>MOD09GA</sub>: PCC calculated for FVC<sub>MOD09GA</sub> and GEOV3 FVC, PCC<sub>MCD43A4</sub>: PCC calculated for FVC<sub>MCD43A4</sub> and GEOV3 FVC. (<b>a</b>) PCC<sub>MOD09GA</sub> and PCC<sub>MCD43A4</sub> in 2018–2021; (<b>b</b>) Scatterplot of numerical differences between PCC<sub>MOD09GA</sub> and PCC<sub>MCD43A4</sub>. SOS: start of season; EOS: end of season; SGOS: stable growth of season.</p>
Full article ">Figure 7
<p>NDVI spatial distribution maps of crop pixel, savanna pixel, and grassland pixel in different stages of the growing seasons. (<b>a</b>–<b>d</b>) Crop. (<b>e</b>–<b>h</b>) Savanna. (<b>i</b>–<b>l</b>) Grassland. SZA: Solar Zenith Angle, FVC: Fractional Vegetation Cover, AFX_RED: Anisotropic Flat Index (AFX) in the red band, AFX_NIR: AFX in the near-infrared band.</p>
Full article ">Figure 8
<p>NDVI spatial distribution maps of mixed pixels in different stages of the growing seasons. (<b>a</b>–<b>d</b>) Crops above and buildings below. (<b>e</b>–<b>h</b>) Crops below and buildings above. (<b>i</b>–<b>l</b>) Buildings in the upper-left corner and crops in the remainder. SZA: Solar Zenith Angle, FVC: Fractional Vegetation Cover, AFX_RED: Anisotropic Flat Index (AFX) in the red band, AFX_NIR: AFX in the near-infrared band.</p>
Full article ">
29 pages, 6780 KiB  
Article
Phenological and Biophysical Mediterranean Orchard Assessment Using Ground-Based Methods and Sentinel 2 Data
by Pierre Rouault, Dominique Courault, Guillaume Pouget, Fabrice Flamain, Papa-Khaly Diop, Véronique Desfonds, Claude Doussan, André Chanzy, Marta Debolini, Matthew McCabe and Raul Lopez-Lozano
Remote Sens. 2024, 16(18), 3393; https://doi.org/10.3390/rs16183393 - 12 Sep 2024
Viewed by 749
Abstract
A range of remote sensing platforms provide high spatial and temporal resolution insights which are useful for monitoring vegetation growth. Very few studies have focused on fruit orchards, largely due to the inherent complexity of their structure. Fruit trees are mixed with inter-rows [...] Read more.
A range of remote sensing platforms provide high spatial and temporal resolution insights which are useful for monitoring vegetation growth. Very few studies have focused on fruit orchards, largely due to the inherent complexity of their structure. Fruit trees are mixed with inter-rows that can be grassed or non-grassed, and there are no standard protocols for ground measurements suitable for the range of crops. The assessment of biophysical variables (BVs) for fruit orchards from optical satellites remains a significant challenge. The objectives of this study are as follows: (1) to address the challenges of extracting and better interpreting biophysical variables from optical data by proposing new ground measurements protocols tailored to various orchards with differing inter-row management practices, (2) to quantify the impact of the inter-row at the Sentinel pixel scale, and (3) to evaluate the potential of Sentinel 2 data on BVs for orchard development monitoring and the detection of key phenological stages, such as the flowering and fruit set stages. Several orchards in two pedo-climatic zones in southeast France were monitored for three years: four apricot and nectarine orchards under different management systems and nine cherry orchards with differing tree densities and inter-row surfaces. We provide the first comparison of three established ground-based methods of assessing BVs in orchards: (1) hemispherical photographs, (2) a ceptometer, and (3) the Viticanopy smartphone app. The major phenological stages, from budburst to fruit growth, were also determined by in situ annotations on the same fields monitored using Viticanopy. In parallel, Sentinel 2 images from the two study sites were processed using a Biophysical Variable Neural Network (BVNET) model to extract the main BVs, including the leaf area index (LAI), fraction of absorbed photosynthetically active radiation (FAPAR), and fraction of green vegetation cover (FCOVER). The temporal dynamics of the normalised FAPAR were analysed, enabling the detection of the fruit set stage. A new aggregative model was applied to data from hemispherical photographs taken under trees and within inter-rows, enabling us to quantify the impact of the inter-row at the Sentinel 2 pixel scale. The resulting value compared to BVs computed from Sentinel 2 gave statistically significant correlations (0.57 for FCOVER and 0.45 for FAPAR, with respective RMSE values of 0.12 and 0.11). Viticanopy appears promising for assessing the PAI (plant area index) and FCOVER for orchards with grassed inter-rows, showing significant correlations with the Sentinel 2 LAI (R2 of 0.72, RMSE 0.41) and FCOVER (R2 0.66 and RMSE 0.08). Overall, our results suggest that Sentinel 2 imagery can support orchard monitoring via indicators of development and inter-row management, offering data that are useful to quantify production and enhance resource management. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>Schematic of the three approaches used to monitor orchard development at different spatial scales throughout the year (from tree level for phenological observations to watershed level using Sentinel 2 data).</p>
Full article ">Figure 2
<p>(<b>a</b>) Locations of the monitored orchards in the Ouvèze–Ventoux watershed (green points at right) and in the La Crau area (yellow points at left), (<b>b</b>) pictures of 2 cherry orchards (13 September and 22 July 2022): top, non-grassed orchard drip-irrigated by two rows of drippers and bottom, grassed orchard drip-irrigated in summer, (<b>c</b>) pictures of 2 orchards in La Crau (top, nectarine tree in spring 22 March 2023 and bottom, in summer 26 June 2022).</p>
Full article ">Figure 3
<p>(<b>a</b>) Main steps in processing the hemispherical photographs. (<b>b</b>) The three methods of data acquisition around the central tree. (<b>c</b>) Protocol used with hemispherical photographs. (<b>d</b>) Protocol used with the Viticanopy application, with 3 trees monitored in the four directions (blue arrows). (<b>e</b>) Protocols used with the ceptometer: P1 measured in the shadow of the trees and (blue) P2 in the inter-rows (black).</p>
Full article ">Figure 4
<p>Protocol for the monitoring of the phenological stages of cherry trees. (<b>a</b>) Phenology of cherry trees according to BBCH; (<b>b</b>) at plot scale, in an orchard, three trees in red monitored by observations (BBCH scale); (<b>c</b>) at tree scale, two locations are selected to classify flowering stage in the tree; and (<b>d</b>) flowering stage of a cherry tree in April 2022.</p>
Full article ">Figure 5
<p>Comparison of temporal profiles of Sentinel 2 LAI interpolated profile (black line) and PAI obtained from the ceptometer (blue line, P2 protocol) and Viticanopy (green line) for three orchards: (<b>a</b>) 3099 (cherry—grassed—Ouvèze), (<b>b</b>) 183 (cherry—non-grassed—Ouvèze), and (<b>c</b>) 4 (nectarine—La Crau) at the beginning of 2023.</p>
Full article ">Figure 6
<p>Comparison between Sentinel 2 LAI and PAI from (<b>a</b>) ceptometer measurements taken at all orchards of the two areas (La Crau and Ouvèze), (<b>b</b>) Viticanopy measurements at all orchards, and (<b>c</b>) Viticanopy measurements excluding 2 non-grassed orchards (183, 259). The black line represents the optimal correlation 1:1; the red line represents the results from linear regression.</p>
Full article ">Figure 7
<p>(<b>a</b>)—(top graphs) Proportion of tree (orange <span class="html-italic">100*FCOVER<sub>t</sub>/FCOVER<sub>c</sub></span>, see Equation (1)) and of inter-row (green <span class="html-italic">100*((1-FCOVER<sub>t</sub>)*FCOVER<sub>g</sub>)/FCOVER<sub>c</sub></span>) components computed from hemispherical photographs used to estimate FCOVER for two dates, 22 March 2022 (doy:81) and 21 June 2022 (doy 172), for all the monitored fields. (<b>b</b>)—(bottom graphs) For two plots, left, field 183.2 and right, field 3099.1, temporal variations in proportion of tree and inter-row components for the different observation dates in 2022.</p>
Full article ">Figure 8
<p>(<b>a</b>) Averaged percentage of grass contribution on FAPAR computed from hemispherical photographs according to Equation (1) for all grassed orchard plots in 2022. Examples of Sentinel 2 FAPAR dynamics (black lines) for plots at (<b>b</b>) non-grassed site 183 and (<b>c</b>) grassed site 1418. Initial values of FAPAR, as computed from BVNET, are provided in black. The green line represents adjusted FAPAR after subtracting the grass contribution (percentage obtained from hemispherical photographs). It corresponds to FAPAR only for the trees. The percentage of grass contribution is in red.</p>
Full article ">Figure 9
<p>Correlation between (<b>a</b>) FCOVER obtained from hemispherical photographs (from Equation (1)) for all orchards of the two studied areas and FCOVER from Sentinel 2 computed with BVNET (<b>b</b>) FAPAR from hemispherical photographs and FAPAR from Sentinel 2 for all orchards and for the 3 years. (<b>c</b>) Correlation between FCOVER from Viticanopy and Sentinel 2 for all orchards for the two areas, except 183 and 259. (<b>d</b>) Correlation between FCOVER from upward-aimed hemispherical photographs and from Viticanopy for all plots.</p>
Full article ">Figure 10
<p>(<b>a</b>) LAI temporal profiles obtained from BVNET applied to Sentinel 2 data averaged at plot and field scales (field 3099) for the year 2022 and (<b>b</b>) soil water stock (in mm in blue) computed at 0–50 cm using capacitive sensors (described in <a href="#sec2dot1-remotesensing-16-03393" class="html-sec">Section 2.1</a>), with rainfall recorded at the Carpentras station (see <a href="#app1-remotesensing-16-03393" class="html-app">Supplementary Part S1 and Table S1</a>).</p>
Full article ">Figure 11
<p>Time series of FCOVER (mean value at field scale) for the cherry trees in field 3099 in Ouvèze area from 2016 to 2023.</p>
Full article ">Figure 12
<p>Sentinel 2 FAPAR evolution in 2022 for two cherry tree fields, with the date of flowering observation (in green) and the date of fruit set observation (in red) for (<b>a</b>) plot 183 (non-grassed cherry trees) and (<b>b</b>) plot 3099 (grassed cherry trees).</p>
Full article ">Figure 13
<p>Variability in dates for the phenological stages of a cherry tree orchard (plot 3099) observed in 2022.</p>
Full article ">Figure 14
<p>(<b>a</b>) Normalised FAPAR computed for all observed cherry trees relative to observation dates for BBCH stages in the Ouvèze area in 2021 for five plots. (<b>b</b>) Map of dates distinguishing between flowering and fruit set stages for 2021 obtained by thresholding FAPAR images.</p>
Full article ">
16 pages, 3639 KiB  
Article
Time-of-Flight Camera Intensity Image Reconstruction Based on an Untrained Convolutional Neural Network
by Tian-Long Wang, Lin Ao, Na Han, Fu Zheng, Yan-Qiu Wang and Zhi-Bin Sun
Photonics 2024, 11(9), 821; https://doi.org/10.3390/photonics11090821 - 30 Aug 2024
Viewed by 829
Abstract
With the continuous development of science and technology, laser ranging technology will become more efficient, convenient, and widespread, and it has been widely used in the fields of medicine, engineering, video games, and three-dimensional imaging. A time-of-flight (ToF) camera is a three-dimensional stereo [...] Read more.
With the continuous development of science and technology, laser ranging technology will become more efficient, convenient, and widespread, and it has been widely used in the fields of medicine, engineering, video games, and three-dimensional imaging. A time-of-flight (ToF) camera is a three-dimensional stereo imaging device with the advantages of small size, small measurement error, and strong anti-interference ability. However, compared to traditional sensors, ToF cameras typically exhibit lower resolution and signal-to-noise ratio due to inevitable noise from multipath interference and mixed pixels during usage. Additionally, in environments with scattering media, the information about objects gets scattered multiple times, making it challenging for ToF cameras to obtain effective object information. To address these issues, we propose a solution that combines ToF cameras with single-pixel imaging theory. Leveraging intensity information acquired by ToF cameras, we apply various reconstruction algorithms to reconstruct the object’s image. Under undersampling conditions, our reconstruction approach yields higher peak signal-to-noise ratio compared to the raw camera image, significantly improving the quality of the target object’s image. Furthermore, when ToF cameras fail in environments with scattering media, our proposed approach successfully reconstructs the object’s image when the camera is imaging through the scattering medium. This experimental demonstration effectively reduces the noise and direct ambient light generated by the ToF camera itself, while opening up the potential application of ToF cameras in challenging environments, such as scattering media or underwater. Full article
Show Figures

Figure 1

Figure 1
<p>Flight time measurement in continuous sinusoidal wave modulation mode.</p>
Full article ">Figure 2
<p>Schematic diagram of the image reconstruction using a neural network. (<b>a</b>) Schematic diagram of network operation, (<b>b</b>) images reconstructed by the neural network with different sampling rates and different number of iterations.</p>
Full article ">Figure 3
<p>The schematic diagrams of SPI.</p>
Full article ">Figure 4
<p>The schematic diagrams of SPI based on a ToF camera.</p>
Full article ">Figure 5
<p>Experimental results of imaging reconstruction using intensity images at different SRs. (<b>a</b>) Target object, (<b>b</b>) ToF image, (<b>c</b>–<b>f</b>) the recovered images by CGI, BP, TVAL3, and DL. The SRs from left to right is 6.25%, 12.5%, 18.75%, 25%, 31.25% and 37.5%.</p>
Full article ">Figure 6
<p>Plots of the PSNRs of the reconstructed intensity images versus the SRs by different algorithms. The black, red, blue, and green lines denote the PSNRs by CGI, BP, TVAL3, and DL.</p>
Full article ">Figure 7
<p>Experimental results of reconstruction using the intensity images through the scattering media at different SRs. (<b>a</b>) ToF image, (<b>b</b>–<b>e</b>) the recovered images by CGI, BP, TVAL3, and DL. The SRs from left to right are 6.25%, 12.5%, 18.75%, 25%, 31.25%, and 37.5%.</p>
Full article ">Figure 8
<p>Plots comparing the PSNR and SRs for the reconstruction of intensity images through scattering media using different algorithms.</p>
Full article ">Figure 9
<p>Experimental results of reconstruction using the intensity images through the scattering media at different SRs. (<b>a</b>) ToF image, (<b>b</b>) ToF image with added Gaussian noise. (<b>c</b>–<b>f</b>) the recovered images by CGI, BP, TVAL3, and DL. The SRs from left to right are 6.25%, 12.5%, 18.75%, 25%, 31.25%, and 37.5%.</p>
Full article ">Figure 10
<p>Plots comparing the PSNR and SRs for the reconstruction of intensity images through scattering media using different algorithms.</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 758
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 ">
27 pages, 79059 KiB  
Article
Unsupervised Noise-Resistant Remote-Sensing Image Change Detection: A Self-Supervised Denoising Network-, FCM_SICM-, and EMD Metric-Based Approach
by Jiangling Xie, Yikun Li, Shuwen Yang and Xiaojun Li
Remote Sens. 2024, 16(17), 3209; https://doi.org/10.3390/rs16173209 - 30 Aug 2024
Viewed by 669
Abstract
The detection of change in remote-sensing images is broadly applicable to many fields. In recent years, both supervised and unsupervised methods have demonstrated excellent capacity to detect changes in high-resolution images. However, most of these methods are sensitive to noise, and their performance [...] Read more.
The detection of change in remote-sensing images is broadly applicable to many fields. In recent years, both supervised and unsupervised methods have demonstrated excellent capacity to detect changes in high-resolution images. However, most of these methods are sensitive to noise, and their performance significantly deteriorates when dealing with remote-sensing images that have been contaminated by mixed random noises. Moreover, supervised methods require that samples are manually labeled for training, which is time-consuming and labor-intensive. This study proposes a new unsupervised change-detection (CD) framework that is resilient to mixed random noise called self-supervised denoising network-based unsupervised change-detection coupling FCM_SICM and EMD (SSDNet-FSE). It consists of two components, namely a denoising module and a CD module. The proposed method first utilizes a self-supervised denoising network with real 3D weight attention mechanisms to reconstruct noisy images. Then, a noise-resistant fuzzy C-means clustering algorithm (FCM_SICM) is used to decompose the mixed pixels of reconstructed images into multiple signal classes by exploiting local spatial information, spectral information, and membership linkage. Next, the noise-resistant Earth mover’s distance (EMD) is used to calculate the distance between signal-class centers and the corresponding fuzzy memberships of bitemporal pixels and generate a map of the magnitude of change. Finally, automatic thresholding is undertaken to binarize the change-magnitude map into the final CD map. The results of experiments conducted on five public datasets prove the superior noise-resistant performance of the proposed method over six state-of-the-art CD competitors and confirm its effectiveness and potential for practical application. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of the proposed SSDNet-FSE framework.</p>
Full article ">Figure 2
<p>Graphical illustration of the SimAM attention mechanism, where the complete 3-D weights are for attention.</p>
Full article ">Figure 3
<p>Network structure of SSDNet.</p>
Full article ">Figure 4
<p>Coupling mechanism of FCM_SICM and EMD.</p>
Full article ">Figure 5
<p>CD results of competitive methods obtained on Shangtang. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 6
<p>CD Results of competitive methods obtained on DSIFN-CD. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth. (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 6 Cont.
<p>CD Results of competitive methods obtained on DSIFN-CD. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth. (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 7
<p>CD results of competitive methods obtained on LZ. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth. (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 8
<p>CD Results of competitive methods obtained on CDD. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth. (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 9
<p>CD Results of competitive methods obtained on GZ. (<b>a</b>) Time 1 image with mixed noises. (<b>b</b>) Time 2 image with mixed noises. (<b>c</b>) Ground truth. (<b>d</b>) GMCD. (<b>e</b>) KPCAMNet. (<b>f</b>) DCVA. (<b>g</b>) PCAKMeans. (<b>h</b>) ASEA. (<b>i</b>) INLPG. (<b>j</b>) Ours.</p>
Full article ">Figure 10
<p>Noise-resistance performance of competitive methods on the five datasets.</p>
Full article ">Figure 11
<p>Change maps obtained by nine ablation methods on GZ dataset.</p>
Full article ">Figure 12
<p>Change-magnitude maps obtained by nine ablation methods on the GZ dataset (real change areas are marked with yellow boundaries).</p>
Full article ">Figure 12 Cont.
<p>Change-magnitude maps obtained by nine ablation methods on the GZ dataset (real change areas are marked with yellow boundaries).</p>
Full article ">Figure 13
<p>Change-magnitude maps obtained by nine ablation methods on the LZ dataset (real change areas are marked with yellow boundaries).</p>
Full article ">Figure 14
<p>Fuzzy level sensitivity on the five datasets.</p>
Full article ">Figure 15
<p>FCM_SICM loss value vs. iteration number.</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 463
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 ">
18 pages, 9929 KiB  
Article
Inversion of Cotton Soil and Plant Analytical Development Based on Unmanned Aerial Vehicle Multispectral Imagery and Mixed Pixel Decomposition
by Bingquan Tian, Hailin Yu, Shuailing Zhang, Xiaoli Wang, Lei Yang, Jingqian Li, Wenhao Cui, Zesheng Wang, Liqun Lu, Yubin Lan and Jing Zhao
Agriculture 2024, 14(9), 1452; https://doi.org/10.3390/agriculture14091452 - 25 Aug 2024
Viewed by 732
Abstract
In order to improve the accuracy of multispectral image inversion of soil and plant analytical development (SPAD) of the cotton canopy, image segmentation methods were utilized to remove the background interference, such as soil and shadow in UAV multispectral images. UAV multispectral images [...] Read more.
In order to improve the accuracy of multispectral image inversion of soil and plant analytical development (SPAD) of the cotton canopy, image segmentation methods were utilized to remove the background interference, such as soil and shadow in UAV multispectral images. UAV multispectral images of cotton bud stage canopies at three different heights (30 m, 50 m, and 80 m) were acquired. Four methods, namely vegetation index thresholding (VIT), supervised classification by support vector machine (SVM), spectral mixture analysis (SMA), and multiple endmember spectral mixture analysis (MESMA), were used to segment cotton, soil, and shadows in the multispectral images of cotton. The segmented UAV multispectral images were used to extract the spectral information of the cotton canopy, and eight vegetation indices were calculated to construct the dataset. Partial least squares regression (PLSR), Random forest (FR), and support vector regression (SVR) algorithms were used to construct the inversion model of cotton SPAD. This study analyzed the effects of different image segmentation methods on the extraction accuracy of spectral information and the accuracy of SPAD modeling in the cotton canopy. The results showed that (1) The accuracy of spectral information extraction can be improved by removing background interference such as soil and shadows using four image segmentation methods. The correlation between the vegetation indices calculated from MESMA segmented images and the SPAD of the cotton canopy was improved the most; (2) At three different flight altitudes, the vegetation indices calculated by the MESMA segmentation method were used as the input variable, and the SVR model had the best accuracy in the inversion of cotton SPAD, with R2 of 0.810, 0.778, and 0.697, respectively; (3) At a flight altitude of 80 m, the R2 of the SVR models constructed using vegetation indices calculated from images segmented by VIT, SVM, SMA, and MESMA methods were improved by 2.2%, 5.8%, 13.7%, and 17.9%, respectively, compared to the original images. Therefore, the MESMA mixed pixel decomposition method can effectively remove soil and shadows in multispectral images, especially to provide a reference for improving the inversion accuracy of crop physiological parameters in low-resolution images with more mixed pixels. Full article
(This article belongs to the Special Issue Application of UAVs in Precision Agriculture—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>Experimental instruments. (<b>a</b>) DJI M300 with a Zenmuse Pl camera, (<b>b</b>) DJI M210 with MS600Pro multispectral camera. Note: The green box in (<b>a</b>) is the Zenmuse P1 camera (DJI, Shenzhen, China), and the red box in (<b>b</b>) is the MS600Pro multispectral camera (Yusense, Inc., Qingdao, China).</p>
Full article ">Figure 3
<p>MESMA under different fertilization gradients. (<b>a1</b>–<b>a3</b>) RGB images, (<b>b1</b>–<b>b3</b>) MNF eigenvalue, (<b>c1</b>–<b>c3</b>) enumerating pixels in an n-dimensional visualizer, (<b>d1</b>–<b>d3</b>) outputting EM spectral.</p>
Full article ">Figure 4
<p>Distribution of pure pixels at different flight altitudes. (<b>a</b>) 30 m; (<b>b</b>) 50 m; (<b>c</b>) 80 m.</p>
Full article ">Figure 5
<p>Segmentation results at different flight altitudes. (<b>a1</b>–<b>a3</b>) RGB images, (<b>b1</b>–<b>b3</b>) <span class="html-italic">NDCSI</span> vegetation index threshold segmentation, (<b>c1</b>–<b>c3</b>) SVM segmentation, (<b>d1</b>–<b>d3</b>) SMA segmentation, (<b>e1</b>–<b>e3</b>) MESMA segmentation.</p>
Full article ">Figure 6
<p>MESMA abundance inversion result map (flight altitude 80 m). (<b>a</b>) cotton; (<b>b</b>) shadow; (<b>c</b>) soil.</p>
Full article ">Figure 7
<p>Correlation between cotton SPAD and vegetation indices at 30 m.</p>
Full article ">Figure 8
<p>Correlation between cotton SPAD and vegetation indices at 50 m.</p>
Full article ">Figure 9
<p>Correlation between cotton SPAD and vegetation indices at 80 m.</p>
Full article ">Figure 10
<p>Inversion results of the optimal cotton SPAD model at different flight altitudes: (<b>a</b>) 30 m; (<b>b</b>) 50 m; (<b>c</b>) 80 m.</p>
Full article ">Figure 11
<p>SPAD distribution map of cotton.</p>
Full article ">
18 pages, 1844 KiB  
Article
PMLPNet: Classifying Multi-Class Pests in Wild Environment via a Novel Convolutional Neural Network
by Liangliang Liu, Jing Chang, Shixin Qiao, Jinpu Xie, Xin Xu and Hongbo Qiao
Agronomy 2024, 14(8), 1729; https://doi.org/10.3390/agronomy14081729 - 6 Aug 2024
Viewed by 697
Abstract
Pest damage is a major factor in reducing crop yield and has negative impacts on the economy. However, the complex background, diversity of pests, and individual differences pose challenges for classification algorithms. In this study, we propose a patch-based neural network (PMLPNet) for [...] Read more.
Pest damage is a major factor in reducing crop yield and has negative impacts on the economy. However, the complex background, diversity of pests, and individual differences pose challenges for classification algorithms. In this study, we propose a patch-based neural network (PMLPNet) for multi-class pest classification. PMLPNet leverages spatial and channel contextual semantic features through meticulously designed token- and channel-mixing MLPs, respectively. This innovative structure enhances the model’s ability to accurately classify complex multi-class pests by providing high-quality local and global pixel semantic features for the fully connected layer and activation function. We constructed a database of 4510 images spanning 40 types of plant pests across 4 crops. Experimental results demonstrate that PMLPNet outperforms existing CNN models, achieving an accuracy of 92.73%. Additionally, heat maps reveal distinctions among different pest images, while patch probability-based visualizations highlight heterogeneity within pest images. Validation on external datasets (IP102 and PlantDoc) confirms the robust generalization performance of PMLPNet. In summary, our research advances intelligent pest classification techniques, effectively identifying various pest types in diverse crop images. Full article
(This article belongs to the Special Issue Advanced Machine Learning in Agriculture)
Show Figures

Figure 1

Figure 1
<p>Example images of the pests. Each image belongs to a different species of pests.</p>
Full article ">Figure 2
<p>The pipeline for the proposed PMLPNet. “T” is transposition operation, which is used to transpose feature maps.</p>
Full article ">Figure 3
<p>The structure for MLP.</p>
Full article ">Figure 4
<p>Confusion matrix thermodynamic image of 40 types of pests on PMLPNet.</p>
Full article ">Figure 5
<p>Qualitative results. (<b>a</b>,<b>c</b>) are original images, and (<b>b</b>,<b>d</b>) are visualized thermal maps.</p>
Full article ">Figure 6
<p>Intra-image heterogeneity by patch predictions. The left image is the original input image. The right image shows the detail of the predicted probability of each patch.</p>
Full article ">Figure 7
<p>Loss and accuracy curves of comparative methods. (<b>a</b>) is the loss cureves of comparative methods, (<b>b</b>) is the accuracy curves of comparative methods.</p>
Full article ">Figure 8
<p>Combination figure of predictive analysis for each class of pests on PMLPNet.</p>
Full article ">
15 pages, 1214 KiB  
Article
A Self-Supervised Few-Shot Semantic Segmentation Method Based on Multi-Task Learning and Dense Attention Computation
by Kai Yi , Weihang Wang  and Yi Zhang 
Sensors 2024, 24(15), 4975; https://doi.org/10.3390/s24154975 - 31 Jul 2024
Viewed by 640
Abstract
Nowadays, autonomous driving technology has become widely prevalent. The intelligent vehicles have been equipped with various sensors (e.g., vision sensors, LiDAR, depth cameras etc.). Among them, the vision systems with tailored semantic segmentation and perception algorithms play critical roles in scene understanding. However, [...] Read more.
Nowadays, autonomous driving technology has become widely prevalent. The intelligent vehicles have been equipped with various sensors (e.g., vision sensors, LiDAR, depth cameras etc.). Among them, the vision systems with tailored semantic segmentation and perception algorithms play critical roles in scene understanding. However, the traditional supervised semantic segmentation needs a large number of pixel-level manual annotations to complete model training. Although few-shot methods reduce the annotation work to some extent, they are still labor intensive. In this paper, a self-supervised few-shot semantic segmentation method based on Multi-task Learning and Dense Attention Computation (dubbed MLDAC) is proposed. The salient part of an image is split into two parts; one of them serves as the support mask for few-shot segmentation, while cross-entropy losses are calculated between the other part and the entire region with the predicted results separately as multi-task learning so as to improve the model’s generalization ability. Swin Transformer is used as our backbone to extract feature maps at different scales. These feature maps are then input to multiple levels of dense attention computation blocks to enhance pixel-level correspondence. The final prediction results are obtained through inter-scale mixing and feature skip connection. The experimental results indicate that MLDAC obtains 55.1% and 26.8% one-shot mIoU self-supervised few-shot segmentation on the PASCAL-5i and COCO-20i datasets, respectively. In addition, it achieves 78.1% on the FSS-1000 few-shot dataset, proving its efficacy. Full article
(This article belongs to the Section Sensing and Imaging)
Show Figures

Figure 1

Figure 1
<p>The overall structure of the proposed self-supervised network. The unsupervised saliency mask is segmented; one part is used for masking and support, and the other part and the entire unsupervised salient area are used to calculate the loss function so as to guide model training.</p>
Full article ">Figure 2
<p>The architecture of our network with the proposed self-supervised meta-learning approach.</p>
Full article ">Figure 3
<p>Illustration of the proposed DACB.</p>
Full article ">Figure 4
<p>Comparison of visualization results on PASCAL-<math display="inline"><semantics> <msup> <mn>5</mn> <mi>i</mi> </msup> </semantics></math>. Columns correspond to the query image with mask, support image with mask, MaskSplit results, and our results.</p>
Full article ">Figure 5
<p>Ablation experiments on the value of parameters a and b.</p>
Full article ">
19 pages, 4490 KiB  
Article
Estimation of Coastal Wetland Vegetation Aboveground Biomass by Integrating UAV and Satellite Remote Sensing Data
by Xiaomeng Niu, Binjie Chen, Weiwei Sun, Tian Feng, Xiaodong Yang, Yangyi Liu, Weiwei Liu and Bolin Fu
Remote Sens. 2024, 16(15), 2760; https://doi.org/10.3390/rs16152760 - 28 Jul 2024
Cited by 2 | Viewed by 1145
Abstract
Aboveground biomass (AGB) serves as a crucial indicator of the carbon sequestration capacity of coastal wetland ecosystems. Conducting extensive field surveys in coastal wetlands is both time-consuming and labor-intensive. Unmanned aerial vehicles (UAVs) and satellite remote sensing have been widely utilized to estimate [...] Read more.
Aboveground biomass (AGB) serves as a crucial indicator of the carbon sequestration capacity of coastal wetland ecosystems. Conducting extensive field surveys in coastal wetlands is both time-consuming and labor-intensive. Unmanned aerial vehicles (UAVs) and satellite remote sensing have been widely utilized to estimate regional AGB. However, the mixed pixel effects in satellite remote sensing hinder the precise estimation of AGB, while high-spatial resolution UAVs face challenges in estimating large-scale AGB. To fill this gap, this study proposed an integrated approach for estimating AGB using field sampling, a UAV, and Sentinel-2 satellite data. Firstly, based on multispectral data from the UAV, vegetation indices were computed and matched with field sampling data to develop the Field–UAV AGB estimation model, yielding AGB results at the UAV scale (1 m). Subsequently, these results were upscaled to the Sentinel-2 satellite scale (10 m). Vegetation indices from Sentinel-2 data were calculated and matched to establish the UAV–Satellite AGB model, enabling the estimation of AGB over large regional areas. Our findings revealed the AGB estimation model achieved an R2 value of 0.58 at the UAV scale and 0.74 at the satellite scale, significantly outperforming direct modeling from field data to satellite (R2 = −0.04). The AGB densities of the wetlands in Xieqian Bay, Meishan Bay, and Hangzhou Bay, Zhejiang Province, were 1440.27 g/m2, 1508.65 g/m2, and 1545.11 g/m2, respectively. The total AGB quantities were estimated to be 30,526.08 t, 34,219.97 t, and 296,382.91 t, respectively. This study underscores the potential of integrating UAV and satellite remote sensing for accurately assessing AGB in large coastal wetland regions, providing valuable support for the conservation and management of coastal wetland ecosystems. Full article
(This article belongs to the Special Issue GIS and Remote Sensing in Ocean and Coastal Ecology)
Show Figures

Figure 1

Figure 1
<p>Study area overview. (<b>a</b>) Study area of Xieqian Bay (XQB). (<b>b</b>) Study area of Meishan Bay (MSB). (<b>c</b>) Study area of Hangzhou Bay (HZB).</p>
Full article ">Figure 2
<p>Photos showing on-site sampling. (<b>a</b>) Real situation of sampling site. (<b>b</b>) Sample setting.</p>
Full article ">Figure 3
<p>Overall workflow of the study (MLR: multiple linear regression; RF: Random Forest).</p>
Full article ">Figure 4
<p>Multivariate linear regression model based on Field–UAV data. (<b>a</b>) Correlation matrix of various vegetation indices from the UAV with wetland vegetation AGB (** indicates <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) Predicted values and measured values of multivariate linear regression based on field and UAV data.</p>
Full article ">Figure 5
<p>RF regression model based on Field–UAV data. (<b>a</b>) Variable importance ranking of UAV vegetation indices in the RF regression model. (<b>b</b>) Changes in the R<sup>2</sup> for models with different numbers of variables. (<b>c</b>) Scatter plot of predicted values and field-observed values for RF regression based on field and UAV data.</p>
Full article ">Figure 6
<p>AGB results estimated by the RF regression model based on Field–UAV data. (<b>a</b>–<b>c</b>) Satellite image and UAV AGB map of FA in XQB, (<b>d</b>–<b>f</b>) satellite image and UAV AGB map of FA in MSB, (<b>g</b>–<b>i</b>) satellite image and UAV AGB map of FA in HZB. (FA: flight area; XQB: Xieqian Bay; MSB: Meishan Bay; HZB: Hangzhou Bay).</p>
Full article ">Figure 7
<p>Multivariate linear regression model based on UAV–Satellite data. (<b>a</b>) Correlation matrix of various vegetation indices from the satellite with wetland vegetation AGB (** indicates <span class="html-italic">p</span> &lt; 0.01). (<b>b</b>) Predicted values and measured values of multivariate linear regression based on UAV and satellite data.</p>
Full article ">Figure 8
<p>RF regression model based on UAV–Satellite data. (<b>a</b>) Variable importance ranking. (<b>b</b>) Variation in the number of variables. (<b>c</b>) Fitting results.</p>
Full article ">Figure 9
<p>AGB estimation results based on satellite data using RF regression model. (<b>a</b>) XQB, (<b>b</b>) MSB, (<b>c</b>) HZB.</p>
Full article ">Figure 10
<p>Fitting performance of AGB estimation models based on Field–Satellite data. (<b>a</b>) MLR, (<b>b</b>) RF regression.</p>
Full article ">
16 pages, 19021 KiB  
Article
MixImages: An Urban Perception AI Method Based on Polarization Multimodalities
by Yan Mo, Wanting Zhou and Wei Chen
Sensors 2024, 24(15), 4893; https://doi.org/10.3390/s24154893 - 28 Jul 2024
Viewed by 555
Abstract
Intelligent urban perception is one of the hot topics. Most previous urban perception models based on semantic segmentation mainly used RGB images as unimodal inputs. However, in natural urban scenes, the interplay of light and shadow often leads to confused RGB features, which [...] Read more.
Intelligent urban perception is one of the hot topics. Most previous urban perception models based on semantic segmentation mainly used RGB images as unimodal inputs. However, in natural urban scenes, the interplay of light and shadow often leads to confused RGB features, which diminish the model’s perception ability. Multimodal polarization data encompass information dimensions beyond RGB, which can enhance the representation of shadow regions, serving as additional data for assistance. Additionally, in recent years, transformers have achieved outstanding performance in visual tasks, and their large, effective receptive field can provide more discriminative cues for shadow regions. For these reasons, this study proposes a novel semantic segmentation model called MixImages, which can combine polarization data for pixel-level perception. We conducted comprehensive experiments on a polarization dataset of urban scenes. The results showed that the proposed MixImages can achieve an accuracy advantage of 3.43% over the control group model using only RGB images in the unimodal benchmark while gaining a performance improvement of 4.29% in the multimodal benchmark. Additionally, to provide a reference for specific downstream tasks, we also tested the impact of different combinations of polarization types on the overall segmentation accuracy. The proposed MixImages can be a new option for conducting urban scene perception tasks. Full article
(This article belongs to the Section Intelligent Sensors)
Show Figures

Figure 1

Figure 1
<p>Overall structure diagram of the proposed MixImages.</p>
Full article ">Figure 2
<p>Overall calculation process of global perception unit.</p>
Full article ">Figure 3
<p>Structure of decoder head.</p>
Full article ">Figure 4
<p>Visualization instances of each modality.</p>
Full article ">Figure 5
<p>Unimodal model qualitative experiments on the RGB-P dataset. Subfigures (<b>a</b>–<b>c</b>) show the experimental results under three different samples. The second to fifth columns of the figures show different colors corresponding to different categories: Background (black), Bicycle (brown), Pedestrian (gray), Sky (cyan), Vegetation (magenta), Road (blue), Car (olive green), Glass (green), and Building (red).</p>
Full article ">Figure 6
<p>Multimodal model qualitative experiments on the RGB-P dataset. Subfigures <b>a</b>–<b>c</b> show the experimental results under three different samples. The correspondence between different colors and categories in columns 2–5 of the figures can be found in the caption of <a href="#sensors-24-04893-f005" class="html-fig">Figure 5</a>.</p>
Full article ">
25 pages, 20390 KiB  
Article
A New and Robust Index for Water Body Extraction from Sentinel-2 Imagery
by Zhenfeng Su, Longwei Xiang, Holger Steffen, Lulu Jia, Fan Deng, Wenliang Wang, Keyu Hu, Jingjing Guo, Aile Nong, Haifu Cui and Peng Gao
Remote Sens. 2024, 16(15), 2749; https://doi.org/10.3390/rs16152749 - 27 Jul 2024
Viewed by 631
Abstract
Land surface water is a key part in the global ecosystem balance and hydrological cycle. Remote sensing has become an effective tool for its spatio-temporal monitoring. However, remote sensing results exemplified in so-called water indices are subject to several limitations. This paper proposes [...] Read more.
Land surface water is a key part in the global ecosystem balance and hydrological cycle. Remote sensing has become an effective tool for its spatio-temporal monitoring. However, remote sensing results exemplified in so-called water indices are subject to several limitations. This paper proposes a new and effective water index called the Sentinel Multi-Band Water Index (SMBWI) to extract water bodies in complex environments from Sentinel-2 satellite imagery. Individual tests explore the effectiveness of the SMBWI in eliminating interference of various special interfering cover features. The Iterative Self-Organizing Data Analysis Technique Algorithm (ISODATA) method and confusion matrix along with the derived accuracy evaluation indicators are used to provide a threshold reference when extracting water bodies and evaluate the accuracy of the water body extraction results, respectively. The SMBWI and eight other commonly used water indices are qualitatively and quantitatively compared through vision and accuracy evaluation indicators, respectively. Here, the SMBWI is proven to be the most effective at suppressing interference of buildings and their shadows, cultivated lands, vegetation, clouds and their shadows, alpine terrain with bare ground and glaciers when extracting water bodies. The overall accuracy in all tests was consistently greater than 96.5%. The SMBWI is proven to have a high ability to identify mixed pixels of water and non-water, with the lowest total error among nine water indices. Most notably, better results are obtained when extracting water bodies under interfering environments of cover features. Therefore, we propose that our novel and robust water index, the SMBWI, is ready to be used for mapping land surface water with high accuracy. Full article
(This article belongs to the Special Issue Remote Sensing for Surface Water Monitoring)
Show Figures

Figure 1

Figure 1
<p>Locations of the six test sites. Test site 1: Dong Lake. Test site 2: Dagang Reservoir. Test site 3: West Lake. Test site 4: Hala Lake. Test site 5: Taiyang Lake. Test site 6: Chengdian Reservoir.</p>
Full article ">Figure 2
<p>Spectral average reflectance(s) of typical land cover features at test site 1, Dong Lake. The horizontal axis are 12 bands of the multi-spectral data from Sentinel-2 satellite image, where B1 represents the coastal aerosol band, B2 blue, B3 green, B4 red, B5, B6, B7, B8 and B8A near-infrared and B9, B11 and B12 shortwave infrared, while the cirrus cloud estimation (B10) bands are not included here.</p>
Full article ">Figure 3
<p>As <a href="#remotesensing-16-02749-f002" class="html-fig">Figure 2</a>, but for spectral average reflectance(s) of typical land cover features at test site 2, Dagang Reservoir.</p>
Full article ">Figure 4
<p>As <a href="#remotesensing-16-02749-f002" class="html-fig">Figure 2</a>, but for spectral average reflectance(s) of typical land cover features at test site 4, Hala Lake.</p>
Full article ">Figure 5
<p>Mixed Pixel Classification at the lakeshore of Chengdian Reservoir. The right side of the image presents a general view of Chengdian Reservoir, while the left side is a magnified section showing the distribution and water content in mixed pixels.</p>
Full article ">Figure 6
<p>Water extraction results of the urban built-up area interference-eliminating test at test site 1, Dong Lake. (<b>a</b>) Original image of the test site; (<b>b</b>–<b>j</b>) extraction results from the corresponding water index. Note that the extracted water bodies are in white, while apart from real water bodies, a fairly substantial portion of the tiny white patches comprises mis-extraction of various other cover features by water indices. Red lines delimit the concentrated mis-extraction patches.</p>
Full article ">Figure 7
<p>Local area extraction results of the urban built-ups’ interference-eliminating test for the best-performing and worst-performing indices. (<b>a</b>) Original image and its local enlarged image. (<b>b</b>–<b>d</b>) Water body extraction results in blue by SMBWI, RNDWI and SWI, respectively, overlying the original image. Red lines delimit the mis-extractions.</p>
Full article ">Figure 8
<p>As <a href="#remotesensing-16-02749-f007" class="html-fig">Figure 7</a>, but for cultivated lands’ interference-eliminating test corresponding to <a href="#remotesensing-16-02749-f006" class="html-fig">Figure 6</a>b,i,j.</p>
Full article ">Figure 9
<p>As <a href="#remotesensing-16-02749-f006" class="html-fig">Figure 6</a>, but for the vegetation and cloud interference-eliminating test at test site 2, Dagang Reservoir.</p>
Full article ">Figure 10
<p>As <a href="#remotesensing-16-02749-f007" class="html-fig">Figure 7</a>, but for the cloud and its shadows’ interference-eliminating test corresponding to <a href="#remotesensing-16-02749-f009" class="html-fig">Figure 9</a>b,i,j.</p>
Full article ">Figure 11
<p>As <a href="#remotesensing-16-02749-f006" class="html-fig">Figure 6</a>, but for the eutrophication and urban built-up shadow test at test site 3, West Lake.</p>
Full article ">Figure 12
<p>As <a href="#remotesensing-16-02749-f011" class="html-fig">Figure 11</a>, but for the magnified display of Beili Lake.</p>
Full article ">Figure 13
<p>As <a href="#remotesensing-16-02749-f007" class="html-fig">Figure 7</a>, but for urban built-up shadow interference-eliminating test corresponding to <a href="#remotesensing-16-02749-f011" class="html-fig">Figure 11</a>b,i,j.</p>
Full article ">Figure 14
<p>As <a href="#remotesensing-16-02749-f006" class="html-fig">Figure 6</a>, but for the eliminating test of alpine terrain with bare ground and glaciers at test site 4, Hala Lake Basin.</p>
Full article ">Figure 15
<p>As <a href="#remotesensing-16-02749-f007" class="html-fig">Figure 7</a>, but for glacier interference-eliminating test corresponding to <a href="#remotesensing-16-02749-f014" class="html-fig">Figure 14</a>b,i,j.</p>
Full article ">Figure 16
<p>Statistical chart of omission errors and commission errors of all nine indices across six water content intervals for the lakeshore mixed pixel test at test site 6, Chengdian Reservoir.</p>
Full article ">Figure A1
<p>Visual interpretation results and water body extraction results for small water bodies on the east of Dagang Reservoir. (<b>a</b>) The selected verification area, with red lines enclosing the small water bodies through visual interpretation. (<b>b</b>–<b>d</b>) the water bodies extracted by SMBWI, MNDWI and RNDWI, respectively, with water bodies in blue, overlying the original imagery.</p>
Full article ">Figure A2
<p>As <a href="#remotesensing-16-02749-f014" class="html-fig">Figure 14</a>, but for the supplementary experiment at test site 5, Taiyang Lake Basin.</p>
Full article ">Figure A3
<p>As <a href="#remotesensing-16-02749-f015" class="html-fig">Figure 15</a>, but for the supplementary experiment at test site 5, Taiyang Lake Basin.</p>
Full article ">
Back to TopTop