[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 (145)

Search Parameters:
Keywords = signal space diversity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 16837 KiB  
Article
MapGen-Diff: An End-to-End Remote Sensing Image to Map Generator via Denoising Diffusion Bridge Model
by Jilong Tian, Jiangjiang Wu, Hao Chen and Mengyu Ma
Remote Sens. 2024, 16(19), 3716; https://doi.org/10.3390/rs16193716 - 6 Oct 2024
Abstract
Online maps are of great importance in modern life, especially in commuting, traveling and urban planning. The accessibility of remote sensing (RS) images has contributed to the widespread practice of generating online maps based on RS images. The previous works leverage an idea [...] Read more.
Online maps are of great importance in modern life, especially in commuting, traveling and urban planning. The accessibility of remote sensing (RS) images has contributed to the widespread practice of generating online maps based on RS images. The previous works leverage an idea of domain mapping to achieve end-to-end remote sensing image-to-map translation (RSMT). Although existing methods are effective and efficient for online map generation, generated online maps still suffer from ground features distortion and boundary inaccuracy to a certain extent. Recently, the emergence of diffusion models has signaled a significant advance in high-fidelity image synthesis. Based on rigorous mathematical theories, denoising diffusion models can offer controllable generation in sampling process, which are very suitable for end-to-end RSMT. Therefore, we design a novel end-to-end diffusion model to generate online maps directly from remote sensing images, called MapGen-Diff. We leverage a strategy inspired by Brownian motion to make a trade-off between the diversity and the accuracy of generation process. Meanwhile, an image compression module is proposed to map the raw images into the latent space for capturing more perception features. In order to enhance the geometric accuracy of ground features, a consistency regularization is designed, which allows the model to generate maps with clearer boundaries and colorization. Compared to several state-of-the-art methods, the proposed MapGen-Diff achieves outstanding performance, especially a 5% RMSE and 7% SSIM improvement on Los Angeles and Toronto datasets. The visualization results also demonstrate more accurate local details and higher quality. Full article
Show Figures

Figure 1

Figure 1
<p>The previous methods for the RSMT have faced some challenges in local detail inaccuracy and semantic distortion. The blue markers show the blurred boundaries of urban roads and buildings; the red rectangular box shows incorrect semantic colorization.</p>
Full article ">Figure 2
<p>Overall architecture of our proposed model MapGen-Diff. Part (<b>a</b>) illustrates the pipeline of MapGen-Diff. There are two diffusion processes with shared weights. First, we use an encoder <span class="html-italic">E</span> to map the remote sensing images from pixel space into feature space. Second, through the transformation function <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>(</mo> <mo>·</mo> <mo>)</mo> </mrow> </semantics></math>=<math display="inline"><semantics> <mrow> <msub> <mi>F</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>F</mi> <mn>2</mn> </msub> <mo>,</mo> <msub> <mi>F</mi> <mn>3</mn> </msub> </mrow> </semantics></math>, we train the network separately on original and transformed data. Finally, we adopt TCR loss to retain the consistency between original data and transformed data. Part (<b>b</b>) shows the detailed diffusion process. We utilize the data distribution from Domain <span class="html-italic">A</span> and Domain <span class="html-italic">B</span> as the endpoints of the diffusion process. While sampling in the reverse process, we utilize a U-Net network in the denoising process to control the generation in the feature space.</p>
Full article ">Figure 3
<p>The details of the mechanism of mapping into latent space. Firstly, when mapping the images <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>∈</mo> <msup> <mi mathvariant="double-struck">R</mi> <mrow> <mi>H</mi> <mo>×</mo> <mi>W</mi> <mo>×</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> into latent space, we use a couple of networks to compress <span class="html-italic">x</span> to a feature vector <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>∈</mo> <msup> <mi mathvariant="double-struck">R</mi> <mrow> <mi>H</mi> <mo>×</mo> <mi>W</mi> <mo>×</mo> <mi>C</mi> </mrow> </msup> </mrow> </semantics></math>. Meanwhile, we propose KL-reg and VQ-reg regularization to avoid high variance, including a reshaping module, calculating module, embedding dictionary and index of the dictionary.</p>
Full article ">Figure 4
<p>Some selected pairs of samples from the two datasets used in this article.</p>
Full article ">Figure 5
<p>The data distributions of the samples in LA datasets and Toronto datasets.</p>
Full article ">Figure 6
<p>Qualitative comparisons of generated online maps by different approaches on LA datasets (red rectangles: out-performance).</p>
Full article ">Figure 7
<p>Qualitative comparisons of generated online maps by different approaches on Toronto datasets (red rectangles: out-performance).</p>
Full article ">Figure 8
<p>Various visualization results of the ablation study with four selected sample in each dataset. The upper three rows are from LA datasets and the ones below are from Toronto datasets.</p>
Full article ">Figure 9
<p>Qualitative results involving varying the maximum variance <math display="inline"><semantics> <mi>δ</mi> </semantics></math>. Part (<b>a</b>) demonstrates the visualization of the generated map; part (<b>b</b>) represents the entropy image of the translated maps based on the different values of <math display="inline"><semantics> <mi>δ</mi> </semantics></math>.</p>
Full article ">Figure 10
<p>Qualitative results with different values of the downsampling factor.</p>
Full article ">Figure 11
<p>Qualitative comparisons with different values of the downsampling factor on two datasets. The bar chart represents the experimental results of diversity. The line chart represents the experimental results of accuracy.</p>
Full article ">Figure 12
<p>The results of different values of sampling steps. Representation of metric values on LA datasets. Representation of metric values on Toronto datasets.</p>
Full article ">
17 pages, 2996 KiB  
Article
Performance Enhancement for B5G/6G Networks Based on Space Time Coding Schemes Assisted by Intelligent Reflecting Surfaces with Higher Modulation Orders
by Mariam El-Hussien, Bassant Abdelhamid, Hesham Elbadawy, Hadia El-Hennawy and Mehaseb Ahmed
Sensors 2024, 24(19), 6169; https://doi.org/10.3390/s24196169 - 24 Sep 2024
Abstract
Intelligent Reflecting Surfaces (IRS) and Multiple-Input Single-Output (MISO) technologies are essential in the fifth generation (5G) networks and beyond. IRS optimizes the signal propagation and the coverage and is a viable approach to address the issues caused by fading channels that limits the [...] Read more.
Intelligent Reflecting Surfaces (IRS) and Multiple-Input Single-Output (MISO) technologies are essential in the fifth generation (5G) networks and beyond. IRS optimizes the signal propagation and the coverage and is a viable approach to address the issues caused by fading channels that limits the spectral efficiency, while MIMO enhances data rates, reliability, and spectral efficiency by using multiple antennas at both transmitter and receiver ends. This paper proposes an IRS-assisted MISO system using the Orthogonal Space-Time Block Code (OSTBC) scheme to enhance the channel reliability and reduce the Bit Error Rate (BER) in wireless communication systems. The proposed system exploits the benefits from the transmit diversity gain of the OSTBC scheme as well as from the bit energy to noise power spectral density (Eb/No) improvement of the IRS technology. The presented work explores these combined technologies across different modulation schemes. The obtained results outperform the similar previously published works by considering higher-order modulation schemes as well as the deployment of rate ¾ OSTBC-assisted IRS. Moreover, the obtained results demonstrate that the integration of OSTBC with IRS can yield significant performance improvements in terms of Eb/No by 7 dB and 13 dB when using 16 reflecting elements and 64 reflecting elements, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>IRS-assisted MISO System Model.</p>
Full article ">Figure 2
<p>Fully Utilized STBC transceiver with IRS.</p>
Full article ">Figure 3
<p>BER versus E<sub>b</sub>/N<sub>o</sub> for the QPSK modulation scheme (<b>a</b>) Alamouti STBC 2 × 1 deployed (<b>b</b>) OSTBC 4 × 1 deployed.</p>
Full article ">Figure 4
<p>BER versus E<sub>b</sub>/N<sub>o</sub> (<b>a</b>) Alamouti STBC employing 16 QAM scheme (<b>b</b>) OSTBC employing 16 QAM scheme (<b>c</b>) Alamouti STBC employing 64 QAM scheme (<b>d</b>) OSTBC employing 64 QAM scheme (<b>e</b>) Alamouti STBC employing 256 QAM scheme (<b>f</b>) OSTBC employing 256 QAM scheme.</p>
Full article ">Figure 4 Cont.
<p>BER versus E<sub>b</sub>/N<sub>o</sub> (<b>a</b>) Alamouti STBC employing 16 QAM scheme (<b>b</b>) OSTBC employing 16 QAM scheme (<b>c</b>) Alamouti STBC employing 64 QAM scheme (<b>d</b>) OSTBC employing 64 QAM scheme (<b>e</b>) Alamouti STBC employing 256 QAM scheme (<b>f</b>) OSTBC employing 256 QAM scheme.</p>
Full article ">Figure 5
<p>BER for Alamouti STBC and OSTBC (4 × 1) using the 16 QAM scheme.</p>
Full article ">Figure 6
<p>BER performance versus the number of reflecting elements at E<sub>b</sub>/N<sub>o</sub> equal to 0 dB.</p>
Full article ">Figure 7
<p>BER performance versus the number of reflecting elements.</p>
Full article ">Figure 8
<p>BER for the QPSK modulation scheme versus the number of IRS reflecting elements (<b>a</b>) Alamouti STBC 2 × 1 deployed (<b>b</b>) OSTBC 4 × 1 deployed at different E<sub>b</sub>/N<sub>o</sub>.</p>
Full article ">Figure 9
<p>BER for the 256 QAM modulation scheme versus the number of IRS reflecting elements (<b>a</b>) Alamouti STBC 2 × 1 deployed (<b>b</b>) OSTBC 4 × 1 deployed at different E<sub>b</sub>/N<sub>o</sub>.</p>
Full article ">
14 pages, 3284 KiB  
Article
Low Complexity Parallel Carrier Frequency Offset Estimation Based on Time-Tagged QPSK Partitioning for Coherent Free-Space Optical Communication
by Siqi Zhang, Liqian Wang, Kunfeng Liu and Shuang Ding
Photonics 2024, 11(9), 885; https://doi.org/10.3390/photonics11090885 - 20 Sep 2024
Abstract
To effectively mitigate the effects of atmospheric turbulence in free space optical (FSO) communication, we propose a parallel carrier frequency offset estimation (FOE) scheme based on time-tagged QPSK partitioning (TTQP). This scheme can be applied to spatial diversity polarization multiplexing (PM) coherent FSO [...] Read more.
To effectively mitigate the effects of atmospheric turbulence in free space optical (FSO) communication, we propose a parallel carrier frequency offset estimation (FOE) scheme based on time-tagged QPSK partitioning (TTQP). This scheme can be applied to spatial diversity polarization multiplexing (PM) coherent FSO communication systems. Specifically, the TTQP scheme performs QPSK partitioning by time-tagging signal points, accurately recording the time intervals between signals, and significantly reducing implementation complexity through a modified Mth power algorithm. The simulation results for the PM 16-quadrature amplitude modulation (QAM) validate the effectiveness of the proposed scheme. Compared to traditional QPSK partitioning algorithms, the TTQP algorithm achieves high accuracy, low complexity, and multi-format versatility in high-speed coherent FSO communication. Full article
(This article belongs to the Special Issue Challenges and Opportunities in Free Space Optical Communication)
Show Figures

Figure 1

Figure 1
<p>Block diagram of digital signal processing (DSP).</p>
Full article ">Figure 2
<p>Principle block diagram of the Time-Tagged QPSK partitioning (TTQP) FOE algorithm.</p>
Full article ">Figure 3
<p>Ideal constellation diagram and actual received signal sequence description for 16QAM. (<b>a</b>) Ideal constellation diagram; (<b>b</b>) Actual received signal Sequence description.</p>
Full article ">Figure 4
<p>The QPSK-TMS signal obtained after the time-tagging QPSK partitioning.</p>
Full article ">Figure 5
<p>The comparison between the results of squaring the sine and cosine functions and those obtained through absolute value operations demonstrates their similarity. (<b>a</b>) Differences between the squared values of the sine function and their absolute values; (<b>b</b>) Differences between the squared values of the cosine function and their absolute values.</p>
Full article ">Figure 6
<p>Calculation and compensation of accumulated frequency offset.</p>
Full article ">Figure 7
<p>Simulation platform for a spatial diversity coherent FSO communication system. Insets: Light field distributions (<b>a</b>) before and (<b>b</b>) and (<b>c</b>) after passing through the phase screen.</p>
Full article ">Figure 8
<p>The relationship between the Normalized Mean Square Error (NMSE) and received optical power at different symbol block lengths under B2B conditions.</p>
Full article ">Figure 9
<p>Performance comparison between the proposed algorithm and traditional QPSK partitioning under B2B conditions: (<b>a</b>) NMSE curve under different average received optical power, (<b>b</b>) BER curve under different average received optical power.</p>
Full article ">Figure 10
<p>Performance comparison between the proposed algorithm and traditional QPSK partitioning under weak turbulence conditions: (<b>a</b>) NMSE curve under different average received optical power, (<b>b</b>) BER curve under different average received optical power.</p>
Full article ">Figure 11
<p>Performance comparison between the proposed algorithm and traditional QPSK Partitioning under strong turbulence conditions: (<b>a</b>) NMSE curve under different average received optical power, (<b>b</b>) BER curve under different average received optical power.</p>
Full article ">Figure 12
<p>Performance comparison between the proposed algorithm and the QPSK algorithm in terms of NMSE under both weak and strong turbulence conditions across different frequency offset ranges: (<b>a</b>): NMSE comparison curve under weak turbulence conditions, (<b>b</b>): NMSE comparison curve under strong turbulence conditions.</p>
Full article ">
24 pages, 2920 KiB  
Article
Opportunistic Interference Alignment in Cognitive Radio Networks with Space–Time Coding
by Yusuf Abdulkadir, Oluyomi Simpson and Yichuang Sun
J. Sens. Actuator Netw. 2024, 13(5), 46; https://doi.org/10.3390/jsan13050046 - 23 Aug 2024
Viewed by 288
Abstract
For a multiuser multiple-input–multiple-output (MIMO) overlay cognitive radio (CR) network, an opportunistic interference alignment (IA) technique has been proposed that allows spectrum sharing between primary users (PUs) and secondary users (SUs) while ensuring zero interference to the PU. The CR system consists of [...] Read more.
For a multiuser multiple-input–multiple-output (MIMO) overlay cognitive radio (CR) network, an opportunistic interference alignment (IA) technique has been proposed that allows spectrum sharing between primary users (PUs) and secondary users (SUs) while ensuring zero interference to the PU. The CR system consists of one PU and K SUs where the PU uses space-time water-filling (ST-WF) algorithm to optimize its transmission and in the process, frees up unused eigenmodes that can be exploited by the SU. The SUs make use of an optimal power allocation algorithm to align their transmitted signals in such a way their interference impairs only the PUs unused eigenmodes. Since the SUs optimal power allocation algorithm turns out to be an optimal beamformer with multiple eigen-beams, this work initially proposes combining the diversity gain property of space-time block codes, the zero-forcing function of IA and beamforming to optimize the SUs transmission rates. This proposed solution requires availability of channel state information (CSI), and to eliminate the need for CSI, this work then combines Differential Space-Time Block Coding (DSTBC) scheme with optimal IA precoders (consisting of beamforming and zero-forcing) to maximize the SUs data rates. Simulation results confirm the accuracy of the proposed solution. Full article
Show Figures

Figure 1

Figure 1
<p>Multiuser CR network model consisting of one PU link and multiple SUs.</p>
Full article ">Figure 2
<p>Average sum rate vs. the SNR at the PUs link for water-filling (SWF and ST-WF) and MEB.</p>
Full article ">Figure 3
<p>Outage probability curves for SWF and ST-WF.</p>
Full article ">Figure 4
<p>(<b>a</b>) Single detection (<b>b</b>) Double detection.</p>
Full article ">Figure 5
<p>Performance comparison of a conventional ED and a double-threshold ED scheme.</p>
Full article ">Figure 6
<p>P<sub>d</sub> vs. SNR with P<sub>f</sub> = <math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.01</mn> </mrow> </semantics></math> using a conventional ED and a double-threshold ED with an <math display="inline"><semantics> <mrow> <mi>R</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> scheme.</p>
Full article ">Figure 7
<p>STBC process.</p>
Full article ">Figure 8
<p>SER Curves for coherent STBC and DSTBC–beamforming schemes.</p>
Full article ">Figure 9
<p>Average sum rate (b/s) against SNR (dB) for two SUs.</p>
Full article ">Figure 10
<p>Average sum rate (b/s) against SNR (dB) for two SUs with DSTBC.</p>
Full article ">
18 pages, 9403 KiB  
Article
Learning-Based Super-Resolution Imaging of Turbulent Flames in Both Time and 3D Space Using Double GAN Architectures
by Chenxu Zheng, Weiming Huang and Wenjiang Xu
Fire 2024, 7(8), 293; https://doi.org/10.3390/fire7080293 - 21 Aug 2024
Viewed by 374
Abstract
This article presents a spatiotemporal super-resolution (SR) reconstruction model for two common flame types, a swirling and then a jet flame, using double generative adversarial network (GAN) architectures. The approach develops two sets of generator and discriminator networks to learn topographic and temporal [...] Read more.
This article presents a spatiotemporal super-resolution (SR) reconstruction model for two common flame types, a swirling and then a jet flame, using double generative adversarial network (GAN) architectures. The approach develops two sets of generator and discriminator networks to learn topographic and temporal features and infer high spatiotemporal resolution turbulent flame structure from supplied low-resolution counterparts at two time points. In this work, numerically simulated 3D turbulent swirling and jet flame structures were used as training data to update the model parameters of the GAN networks. The effectiveness of our model was then thoroughly evaluated in comparison to other traditional interpolation methods. An upscaling factor of 2 in space, which corresponded to an 8-fold increase in the total voxel number and a double time frame acceleration, was used to verify the model’s ability on a swirling flame. The results demonstrate that the assessment metrics, peak signal-to-noise ratio (PSNR), overall error (ER), and structural similarity index (SSIM), with average values of 35.27 dB, 1.7%, and 0.985, respectively, in the spatiotemporal SR results, can reach acceptable accuracy. As a second verification to highlight the present model’s potential universal applicability to flame data of diverse types and shapes, we applied the model to a turbulent jet flame and had equal success. This work provides a different method for acquiring high-resolution 3D structure and further boosting repeat rate, demonstrating the potential of deep learning technology for combustion diagnosis. Full article
(This article belongs to the Special Issue Combustion Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Presentation of simulation data: (<b>a</b>–<b>c</b>) isosurface rendering of 3D swirl flame; (<b>d</b>–<b>f</b>) 2D central slice.</p>
Full article ">Figure 2
<p>Mean and RMS velocity comparison of LES simulation in this work and experimental data [<a href="#B39-fire-07-00293" class="html-bibr">39</a>] of the swirl flame at different stations at the indicated distance from the bluff-body.</p>
Full article ">Figure 3
<p>Architecture of spatiotemporal super-resolution network for 3D flame reconstruction based on GAN.</p>
Full article ">Figure 4
<p>Evolution of the loss functions for the training process: (<b>a</b>) loss function variation of spatial SR training and (<b>b</b>) loss function variation of temporal SR training.</p>
Full article ">Figure 5
<p>The quantitative comparison between our spatial SR model and three traditional interpolation methods, namely, nearest, linear, and cubic: (<b>a</b>) variation of PSNR; (<b>b</b>) variation of <span class="html-italic">E<sub>R</sub></span>; (<b>c</b>) variation of SSIM.</p>
Full article ">Figure 6
<p>The comparison of temporal SR results of the SR model and linear interpolation. (<b>a</b>) Slicing images of the swirling flame at two time frames (T<sub>0</sub>, T<sub>1</sub>) and the SR images of time based on the SR model and linear interpolation. (<b>b</b>) One-dimensional comparison of the results at the red dashed line.</p>
Full article ">Figure 7
<p>The quantitative comparison between temporal inbetweening and the result of three interpolation methods: (<b>a</b>) variation of PSNR; (<b>b</b>) variation of <span class="html-italic">E<sub>R</sub></span>; (<b>c</b>) variation of SSIM.</p>
Full article ">Figure 8
<p>Visual comparison of spatiotemporal SR network and cubic interpolation: (<b>a</b>–<b>d</b>) 3D structure topography and (<b>e</b>–<b>h</b>) 2D central slice.</p>
Full article ">Figure 9
<p>Zoomed illustration of a local area and intensity variation with the inbetweening results of different methods. (<b>e</b>) One-dimensional comparison of the results at the <span class="html-italic">X</span> = 10 red dashed line. (<b>j</b>) One-dimensional comparison of the results at the <span class="html-italic">X</span> = 20 red dashed line.</p>
Full article ">Figure 10
<p>Degradation of SR quality due to salt-and-pepper noise added. (<b>a</b>) variation of PSNR; (<b>b</b>) variation of <span class="html-italic">E<sub>R</sub></span>; (<b>c</b>) variation of SSIM.</p>
Full article ">Figure 11
<p>The performance of the 3D spatiotemporal reconstruction model for jet flame: (<b>a</b>–<b>c</b>) the immediate results of spatial SR; (<b>d</b>–<b>f</b>) the ultimate results of spatiotemporal SR.</p>
Full article ">Figure 12
<p>Visual comparison of the spatiotemporal SR reconstruction of jet flame: (<b>a</b>–<b>d</b>) 3D structure; (<b>e</b>–<b>h</b>) 2D central slice.</p>
Full article ">Figure 13
<p>Local amplification and intensity variation comparison of the spatiotemporal SR results of jet flame. (<b>e</b>) One-dimensional comparison of the results at the <span class="html-italic">Z</span> = 35. (<b>j</b>) One-dimensional comparison of the results at the <span class="html-italic">Z</span> = 70.</p>
Full article ">Figure 14
<p>Validation of noise immunity for the pre-trained model by jet flame. (<b>a</b>) variation of PSNR; (<b>b</b>) variation of <span class="html-italic">E<sub>R</sub></span>; (<b>c</b>) variation of SSIM.</p>
Full article ">
29 pages, 13796 KiB  
Article
Clutter Rank Estimation Method for Bistatic Radar Systems Based on Prolate Spheroidal Wave Functions
by Xiao Tan, Zhiwei Yang, Xianghai Li, Lei Liu and Xiaorui Li
Remote Sens. 2024, 16(16), 2928; https://doi.org/10.3390/rs16162928 - 9 Aug 2024
Viewed by 736
Abstract
Bistatic radar exhibits spatial isomerism and diverse configurations, leading to unique clutter characteristics distinct from those of monostatic radar. The clutter rank serves as a pivotal indicator of clutter characteristics, enabling the quantification of clutter severity. Space-time adaptive processing (STAP) is a critical [...] Read more.
Bistatic radar exhibits spatial isomerism and diverse configurations, leading to unique clutter characteristics distinct from those of monostatic radar. The clutter rank serves as a pivotal indicator of clutter characteristics, enabling the quantification of clutter severity. Space-time adaptive processing (STAP) is a critical technique to detect moving targets, and clutter rank determines the number of independent and identically distributed (IID) training samples and the degree of freedom (DOF) for effective suppression of clutter that STAP requires. Therefore, the accurate estimation of clutter rank for bistatic radar can provide a crucial indicator for designing and constructing STAP processors, thereby facilitating fast and efficient clutter suppression in bistatic radar systems. This study is based on the idea that clutter rank is the number of prolate spheroidal wave function (PSWF) orthogonal bases utilized for approximating the clutter signal. Firstly, the challenge of utilizing PSWF orthogonal bases for approximating the clutter signal in bistatic radar is elucidated. This pertains to the fact that, unlike monostatic radar clutter, bistatic radar clutter is not capable of being expressed as a single-frequency signal. The clutter rank estimation for bistatic radar is thus derived as the frequency bandwidth estimation. Secondly, to achieve this estimation, the frequency distribution of each individual scattering unit is investigated, thereby determining their extending frequency broadening (EFB) as compared to that of single-frequency. Subsequently, the integral average of EFB across the entire range bin is computed, ultimately enabling the acquisition of bistatic radar’s frequency bandwidth. Finally, the estimation method is extended to non-side-looking mode and limited observation areas with pattern modulation. Simulation experiments confirm that our proposed method provides accurate clutter rank estimations, surpassing 99% proportions of large eigenvalues across various bistatic configurations, observation modes, and areas. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the receiving platform velocity coordinate system (RVCS).</p>
Full article ">Figure 2
<p>Schematic of the effects of Earth’s rotation are transferred to the satellite movements.</p>
Full article ">Figure 3
<p>Extended frequency values in space–time steering against the scattering units in a typical bistatic radar configuration. (<b>a</b>) Range of the extend frequency values surpasses that of <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>s</mi> <mo>,</mo> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>b</b>) Range of the extended frequency values is smaller than that of <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>s</mi> <mo>,</mo> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Variation schematic of the CNR after beamforming within the coverage area under pattern modulation.</p>
Full article ">Figure 5
<p>Clutter rank versus the CNR after beamforming of the far side lobe in a typical bistatic radar configuration.</p>
Full article ">Figure 6
<p>Schematic of the frequency values distribution in the limited observation area.</p>
Full article ">Figure 7
<p>Distribution of <math display="inline"><semantics> <mrow> <msub> <mi>f</mi> <mrow> <mi>s</mi> <mi>i</mi> </mrow> </msub> <mo>+</mo> <mo>Δ</mo> <msub> <mi>f</mi> <mrow> <mi>i</mi> <mo>,</mo> <mi>max</mi> </mrow> </msub> </mrow> </semantics></math> in the limited area with poor range solution.</p>
Full article ">Figure 8
<p>Schematic of the relative coordinate (RC) system.</p>
Full article ">Figure 9
<p>Schematic of relationship between the transmitting and receiving aircraft velocities in the xoz plane. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Statistical results and proportions of the clutter rank estimation versus <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>. (<b>a</b>) The statistical results. (<b>b</b>) The proportions.</p>
Full article ">Figure 11
<p>Statistical results and proportions of the clutter rank estimation versus <span class="html-italic">δ<sub>tα</sub></span> with <span class="html-italic">δ<sub>rα</sub></span> = −30°. (<b>a</b>) The statistical results. (<b>b</b>) The proportions.</p>
Full article ">Figure 12
<p>Distribution of the eigenvalue with <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>10</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 13
<p>Statistical results and proportions of the clutter rank estimation versus <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> with <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math> under different aircraft altitudes. (<b>a</b>) The statistical results. (<b>b</b>) The proportions.</p>
Full article ">Figure 14
<p>Schematic of relationship between the transmitting and receiving satellite velocities in the xoz plane. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>Distribution curves of the extended frequency value in various LEO-LEO bistatic configurations. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>. (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>. (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Distribution curves of the extended frequency value of <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math> in (<b>a</b>) MEO-LEO and (<b>b</b>) GEO-LEO bistatic configurations.</p>
Full article ">Figure 17
<p>Distribution of the eigenvalue in various bistatic configurations. (<b>a</b>) LEO-LEO, (<b>b</b>) MEO-LEO, and (<b>c</b>) GEO-LEO.</p>
Full article ">Figure 18
<p>Distribution of the eigenvalue with various yaw angles <math display="inline"><semantics> <mrow> <msub> <mo>Δ</mo> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> under LEO-LEO bistatic configurations of (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 18 Cont.
<p>Distribution of the eigenvalue with various yaw angles <math display="inline"><semantics> <mrow> <msub> <mo>Δ</mo> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> under LEO-LEO bistatic configurations of (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <msup> <mrow> <mn>30</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 19
<p>Statistical results and proportions of the clutter rank estimation versus <math display="inline"><semantics> <mrow> <msub> <mo>Δ</mo> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> of various LEO-LEO bistatic configurations. (<b>a</b>) The estimated clutter rank. (<b>b</b>) The proportion.</p>
Full article ">Figure 20
<p>Clutter rank estimation results with <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>90</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, respectively, in (<b>a</b>–<b>d</b>) when the velocity directions of the transmitting and receiving platforms are parallel to the ground. <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> in (<b>a</b>–<b>d</b>) correspond to <math display="inline"><semantics> <mrow> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>90</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mo>−</mo> <msup> <mrow> <mn>45</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mo>−</mo> <msup> <mrow> <mn>90</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>−</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 21
<p>Schematic of observation area 1 to 4 according to the azimuth resolution distribution in the xoz plane. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <mo> </mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 22
<p>Distribution of the eigenvalue in observation area 1 to 4. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mrow> <mn>135</mn> </mrow> <mo>°</mo> </msup> <mo>,</mo> <mo> </mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 23
<p>Statistical results of the clutter rank estimation versus <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>β</mi> </mrow> </msub> </mrow> </semantics></math> when <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> remain constant. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> are 0°, 15°, 30°, 45°, 60°, 75° and 90°. (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> are 0°, −15°, −30°, −45°, −60°, −75° and −90°.</p>
Full article ">Figure 24
<p>Proportions of clutter rank estimation versus <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>β</mi> </mrow> </msub> </mrow> </semantics></math> when <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> remain constant. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> are 0°, 15°, 30°, 45°, 60°, 75° and 90°. (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> </mrow> </msub> </mrow> </semantics></math> are 0°, −15°, −30°, −45°, −60°, −75° and −90°.</p>
Full article ">Figure A1
<p>Illustration of extended frequency value in the space–time steering distribution with <math display="inline"><semantics> <mrow> <msub> <mi>δ</mi> <mrow> <mi>t</mi> <mi>α</mi> <mn>0</mn> </mrow> </msub> <mo>=</mo> <msub> <mi>δ</mi> <mrow> <mi>r</mi> <mi>α</mi> </mrow> </msub> <mo>=</mo> <msup> <mn>0</mn> <mo>°</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">
22 pages, 16238 KiB  
Article
Spectroscopic Phenological Characterization of Mangrove Communities
by Christopher Small and Daniel Sousa
Remote Sens. 2024, 16(15), 2796; https://doi.org/10.3390/rs16152796 - 30 Jul 2024
Viewed by 473
Abstract
Spaceborne spectroscopic imaging offers the potential to improve our understanding of biodiversity and ecosystem services, particularly for challenging and rich environments like mangroves. Understanding the signals present in large volumes of high-dimensional spectroscopic observations of vegetation communities requires the characterization of seasonal phenology [...] Read more.
Spaceborne spectroscopic imaging offers the potential to improve our understanding of biodiversity and ecosystem services, particularly for challenging and rich environments like mangroves. Understanding the signals present in large volumes of high-dimensional spectroscopic observations of vegetation communities requires the characterization of seasonal phenology and response to environmental conditions. This analysis leverages both spectroscopic and phenological information to characterize vegetation communities in the Sundarban riverine mangrove forest of the Ganges–Brahmaputra delta. Parallel analyses of surface reflectance spectra from NASA’s EMIT imaging spectrometer and MODIS vegetation abundance time series (2000–2022) reveal the spectroscopic and phenological diversity of the Sundarban mangrove communities. A comparison of spectral and temporal feature spaces rendered with low-order principal components and 3D embeddings from Uniform Manifold Approximation and Projection (UMAP) reveals similar structures with multiple spectral and temporal endmembers and multiple internal amplitude continua for both EMIT reflectance and MODIS Enhanced Vegetation Index (EVI) phenology. The spectral and temporal feature spaces of the Sundarban represent independent observations sharing a common structure that is driven by the physical processes controlling tree canopy spectral properties and their temporal evolution. Spectral and phenological endmembers reside at the peripheries of the mangrove forest with multiple outward gradients in amplitude of reflectance and phenology within the forest. Longitudinal gradients of both phenology and reflectance amplitude coincide with LiDAR-derived gradients in tree canopy height and sub-canopy ground elevation, suggesting the influence of surface hydrology and sediment deposition. RGB composite maps of both linear (PC) and nonlinear (UMAP) 3D feature spaces reveal a strong contrast between the phenological and spectroscopic diversity of the eastern Sundarban and the less diverse western Sundarban. Full article
(This article belongs to the Special Issue Remote Sensing of Land Surface Phenology II)
Show Figures

Figure 1

Figure 1
<p>Index map of the Sundarban mangrove forest at the mouths of the Ganges–Brahmaputra delta. The Sentinel 2 false color composite from 2018 shows river channel network and forest canopy cover variations. Note the contrast of the mangrove canopy with the dry season agriculture (bright green), fallow fields (tan), and aquaculture ponds (black) on embanked islands surrounding the forest. The contrast in mangrove reflectance between the eastern and western tiles is a BRDF effect due to the contrasting view geometries at the opposite edges of adjacent Sentinel 2 swaths. GPS tracks (white) show the extents of boat-based field surveys. The east–west scale is 185 km.</p>
Full article ">Figure 2
<p>Sundarban EMIT mosaic. Three swaths provide a full coverage of mangroves and surrounding agriculture and aquaculture, with significant swath overlap. Acquisition dates span nearly the full annual phenological cycle from post-monsoon in December to pre-monsoon in April. Solar zenith angles at times of acquisition range from 11.5° (04.24) to 50.5° (12.27). The white vector boundary shows the extent of the Bangladesh Sundarban, for which tree species maps are available. Acquisition times are UTC + 6 h offset.</p>
Full article ">Figure 3
<p>Spectral feature space of the vegetation-masked Sundarban EMIT mosaic. Orthogonal projections of three low-order principal components (PCs) reveal spectral endmembers (labeled) and distinct amplitude gradients (vectors) within three clusters. Varying amplitude reflectance spectra (right) correspond to vector continua of the same color (left). While the reflectance spectra of the clusters overlap at VNIR wavelengths, each is distinct in the SWIR. The feature space spans amplitude continua with both agricultural (Ag) and forest (F) components. The agricultural continuum arises from the varying abundance of photosynthetic (PV) and non-photosynthetic (NPV) vegetation, but the amplitude of mangrove reflectance is modulated primarily by the canopy structure and varying amounts of crown shadow. There is a considerable overlap in EVI range between adjacent gradients, but a negligible overlap in NDWI range.</p>
Full article ">Figure 4
<p>Complementary spectral feature spaces of the vegetation-masked Sundarban EMIT mosaic. Orthogonal projections of two 3D UMAP embeddings (nn: 50 and 100) reveal consistent spectral endmembers (labeled) and distinct reflectance amplitude gradients (vectors) within three clusters corresponding to those in <a href="#remotesensing-16-02796-f003" class="html-fig">Figure 3</a>. In both UMAP 3/2 projections (right), the western continuum (yellow) is distinct from the connected eastern and central continua (cyan and magenta). Note the bifurcation of the high-amplitude end of the eastern continuum into the northern (N) and southern (S) peripheries of the mangroves. As in the PC feature space, the surrounding agriculture forms a separate 2D continuum spanning photosynthetic and non-photosynthetic vegetation.</p>
Full article ">Figure 5
<p>PC and UMAP feature space composites. Reflectance amplitude gradients and inset spectra correspond to those in <a href="#remotesensing-16-02796-f003" class="html-fig">Figure 3</a> and <a href="#remotesensing-16-02796-f004" class="html-fig">Figure 4</a>. Gradient vectors indicate a direction of increasing NIR reflectance. Composite colors are determined by 3D feature space topology in PC and UMAP spaces. The swath edge discontinuity in the center contrasts post-monsoon (west) from dry season (east) reflectance and longitudinal gradients in species composition and environmental conditions.</p>
Full article ">Figure 6
<p>Bitemporal reflectance change in swath overlap. PCs of the bitemporal reflectance space show a continuum bounded by three endmember reflectance changes for mangrove forest and one for dry season agriculture (Ag). For each forest change endmember, SWIR liquid water absorptions are deeper in December, following the monsoon, but significantly reduced by the April dry season. In contrast, the chlorophyll absorptions in the visible change little. Coherent spatial patterns in bitemporal PC composite suggest aggregate responses to solar illumination (θ) and SWIR water absorption.</p>
Full article ">Figure 7
<p>Temporal feature space of MODIS EVI time series for the entire Sundarban mangrove forest, 2000–2022. The UMAP composite (<b>upper left</b>) shows both N–S and E–W gradients in seasonal phenology, as well as several abrupt transitions. The 1/3 projection of the 3D UMAP feature space (<b>upper center</b>) has a single root (11) corresponding to lower EVI mixtures of canopy, water, and shadow at riverbanks and narrow channels within the mangrove. As EVI increases with canopy closure, the root diverges (10) into mixing trends, terminating at 9 distinct temporal endmembers. EVI time series (<b>bottom</b>) increase abruptly during the summer monsoon, and then decrease gradually over the rest of the year. The 9 endmembers correspond to peripheral regions of the Sundarban (<b>upper left</b>) with the highest post-monsoon EVI. The correlation matrix (<b>upper right</b>) of all 11 endmembers shows the highest correlations between geographically adjacent endmembers. The 2 lowest- (10, 11) and 2 highest (1, 9)-amplitude endmembers are less intercorrelated than those at the southern and eastern peripheries (2–8). Also apparent in the UMAP composite is the distinction between the phenological diversity of the eastern Sundarban and the more homogeneous center in the west.</p>
Full article ">Figure 8
<p>Sundarban temporal endmember phenologies from the temporal feature space in <a href="#remotesensing-16-02796-f007" class="html-fig">Figure 7</a>. The mean EVI (white) shows the rapid post-monsoon greening and gradual dry season senescence, while mean-removed residuals (color) of individual endmembers (offset for clarity) show a diversity of periodic excursions from the mean. As seen in <a href="#remotesensing-16-02796-f007" class="html-fig">Figure 7</a>, all endmembers are phase-aligned and differ primarily in the rate and amplitude of dry season EVI decrease. Despite the considerable noise, distinct annual periodicity is apparent in all but the lowest-amplitude (e.g., 2, 4, 7) residuals—which are most similar to the mean. The largest-amplitude residuals are those from the root and branch (10, 11) of the feature space, corresponding to lower EVI associated with partial canopy cover on shorelines and small channels. Note the slight decadal increase in minimum EVI of the mean.</p>
Full article ">Figure 9
<p>EMIT reflectance mosaic UMAP composite and elevation maps for the Sundarban and surrounding delta. GEDI LiDAR maps (center) reveal that an ongoing sediment deposition in the mangrove forest results in 1–2 m higher ground elevation in the eastern Sundarban relative to the surrounding embanked islands, which have been sediment-starved for decades. The higher SRTM elevation of the eastern Sundarban is a result of both the higher ground elevation and the greater canopy height of the tree species. Mono-species epicenters, from a Bangladesh Forest Department species map (2002), are labeled by common (local) names of tree species. Bi-species gradients compose most of the eastern Sundarban. Arrows show two SRTM swath discontinuity artifacts, which are distinct from the numerous height discontinuities occurring across channels.</p>
Full article ">Figure 10
<p>Field photos illustrate the forest diversity of the Bangladesh Sundarban. The northeast Sundarban (<b>top</b>) reaches canopy heights of 25 m, in contrast with the surrounding embanked islands, which are often below sea level. The sand-dominant islands of the southeast (<b>upper center</b>) are intertidal only around their peripheries and contain different tree species from the rest of the Sundarban. The vegetation gradient of Bird Island on the Bay of Bengal (<b>lower center</b>) illustrates the succession of grasses, shrubs, and trees that colonize sand-dominant islands. River channel networks (<b>bottom</b>) continually deliver silt and mud to intertidal islands throughout the Sundarban. Photos © C. Small 2012–2022.</p>
Full article ">Figure A1
<p>Multiscale UMAP temporal feature space with 3D PC(UMAP<sub>10+50</sub>) composite for MODIS EVI phenology. The low-order PCs of two 3D UMAP embeddings with contrasting n neighbor scales (nn: 10 and 50) preserve both the global scale limb structure and the finer scale clusters that are both phenologically and geographically distinct—including anomalous tree species assemblages at Hiron Point (HP) and Shelar Char (SC) on the Bay of Bengal shorelines. Compare the map structure with the maps in <a href="#remotesensing-16-02796-f005" class="html-fig">Figure 5</a> and <a href="#remotesensing-16-02796-f007" class="html-fig">Figure 7</a>. Manifold density and UMAP color scale equivalent to those in <a href="#remotesensing-16-02796-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure A2
<p>Variance partition and sparse component distribution for the MODIS EVI phenology of the Ganges–Brahmaputra delta. The singular values (top) of the low-rank component suggest that the temporal feature space is effectively 4D (&gt;1%) with 96% of the total variance, while the sparse component has a nearly uniform noise floor over all dimensions. The spatial standard deviation (σ) and range (ρ) of the sparse component peak during the monsoon as a result of a transient cloud cover.</p>
Full article ">Figure A3
<p>Coregistered overlap between 12.23 and 12.27 EMIT acquisitions. Natural color composites illustrate the difference in aerosol optical depth with a reduced dynamic range and a greater adjacency effect on 12.27.</p>
Full article ">Figure A4
<p>Coregistered overlap between 12.27 and 04.24 acquisitions. Natural color composites show the difference in aerosol optical depth with reduced dynamic range and greater adjacency effect on 12.27.</p>
Full article ">Figure A5
<p>Apparent change in reflectance for overlaps on 12.23, 12.27, and 04.24. The mean (white) ± 1 standard deviation (green) of all vegetation spectra in each swath overlap show the effects of residual atmospheric scattering on 12.27 and actual changes in illumination and leaf water content on 04.24. The December 23 and 27 difference suggests short wavelength-dependent scattering on 12.27 with increased visible and reduced NIR but negligible change in SWIR wavelengths. In contrast, a greater VNIR scatter on 12.27 is manifested as a reduced visible and increased NIR scatter relative to 04.24. The reduced leaf water content on 04.24 results in greater SWIR residual from reduced H<sub>2</sub>O absorption after the 4-month dry season. A higher solar elevation in April also contributes to higher NIR and SWIR reflectance.</p>
Full article ">
22 pages, 6002 KiB  
Article
Latent Prototype-Based Clustering: A Novel Exploratory Electroencephalography Analysis Approach
by Sun Zhou, Pengyi Zhang and Huazhen Chen
Sensors 2024, 24(15), 4920; https://doi.org/10.3390/s24154920 - 29 Jul 2024
Viewed by 453
Abstract
Electroencephalography (EEG)-based applications in brain–computer interfaces (BCIs), neurological disease diagnosis, rehabilitation, etc., rely on supervised approaches such as classification that requires given labels. However, with the ever-increasing amount of EEG data, incomplete or incorrectly labeled or unlabeled EEG data are increasing. It likely [...] Read more.
Electroencephalography (EEG)-based applications in brain–computer interfaces (BCIs), neurological disease diagnosis, rehabilitation, etc., rely on supervised approaches such as classification that requires given labels. However, with the ever-increasing amount of EEG data, incomplete or incorrectly labeled or unlabeled EEG data are increasing. It likely degrades the performance of supervised approaches. In this work, we put forward a novel unsupervised exploratory EEG analysis solution by clustering based on low-dimensional prototypes in latent space that are associated with the respective clusters. Having the prototype as a baseline of each cluster, a compositive similarity is defined to act as the critic function in clustering, which incorporates similarities on three levels. The approach is implemented with a Generative Adversarial Network (GAN), termed W-SLOGAN, by extending the Stein Latent Optimization for GANs (SLOGAN). The Gaussian Mixture Model (GMM) is utilized as the latent distribution to adapt to the diversity of EEG signal patterns. The W-SLOGAN ensures that images generated from each Gaussian component belong to the associated cluster. The adaptively learned Gaussian mixing coefficients make the model remain effective in dealing with an imbalanced dataset. By applying the proposed approach to two public EEG or intracranial EEG (iEEG) epilepsy datasets, our experiments demonstrate that the clustering results are close to the classification of the data. Moreover, we present several findings that were discovered by intra-class clustering and cross-analysis of clustering and classification. They show that the approach is attractive in practice in the diagnosis of the epileptic subtype, multiple labelling of EEG data, etc. Full article
(This article belongs to the Special Issue EEG Signal Processing Techniques and Applications—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Different periods of electroencephalography (EEG) signals of an epileptic. a–e denote different time points.</p>
Full article ">Figure 2
<p>Schematic of EEG clustering solution based on latent prototypes. CWT, continuous wavelet transform. DFM, deep feature map. <b><span class="html-italic">e</span></b><sub>query</sub>, latent space representation of the query signal. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mi mathvariant="bold-italic">k</mi> </mrow> </msub> </mrow> </semantics></math>, latent prototype of the <span class="html-italic">k</span>th cluster. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mi mathvariant="normal">q</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">y</mi> </mrow> </msub> </mrow> </semantics></math>, scalogram of the query signal. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>x</mi> </mrow> <mrow> <mi>k</mi> </mrow> </msub> </mrow> </semantics></math>, baseline scalogram of the <span class="html-italic">k</span>th cluster. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mi mathvariant="normal">q</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">y</mi> </mrow> </msub> </mrow> </semantics></math><sub>,</sub> deep feature map of the query signal. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mi>k</mi> </mrow> </msub> </mrow> </semantics></math>, baseline deep feature map of the <span class="html-italic">k</span>th cluster. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>α</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> are weights.</p>
Full article ">Figure 3
<p>Latent distribution defined as Gaussian mixture distribution and distribution of generated data and that of real data. Suppose there are three clusters in the dataset. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> can be regarded as the latent prototypes of the three clusters.</p>
Full article ">Figure 4
<p>Network architecture of W-SLOGAN. The latent distribution is defined as Gaussian mixture distribution. Assume the number of Gaussian components is 3. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">z</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">z</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">z</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the latent vectors sampled from latent space. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the encoded vectors of the scalograms calculated by the encoder. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the mean vectors of the three Gaussian components, corresponding to the latent prototypes of the three clusters. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> denotes the output of the discriminator.</p>
Full article ">Figure 5
<p>Three levels of similarity for clustering. Assume the number of Gaussian components is 3. DFM: deep feature map. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> <mo>,</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">μ</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the mean vectors of the three Gaussian components, corresponding to the latent prototypes of the three clusters. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">e</mi> </mrow> <mrow> <mi mathvariant="normal">q</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">y</mi> </mrow> </msub> </mrow> </semantics></math> denotes the latent representation of the query signal. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">x</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">x</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">x</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the baseline scalograms of the three clusters. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="bold-italic">x</mi> </mrow> <mrow> <mi mathvariant="normal">q</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">y</mi> </mrow> </msub> </mrow> </semantics></math> denotes the scalogram of the query signal. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mn>3</mn> </mrow> </msub> </mrow> </semantics></math> denote the baseline deep feature maps of the three clusters. <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>D</mi> <mi>F</mi> <mi>M</mi> </mrow> <mrow> <mi mathvariant="normal">q</mi> <mi mathvariant="normal">u</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">r</mi> <mi mathvariant="normal">y</mi> </mrow> </msub> </mrow> </semantics></math> denotes the deep feature map of the query signal.</p>
Full article ">Figure 6
<p>Clustering results and intra-class diversity. (<b>A1</b>–<b>A3</b>) show the probability density functions for samples belonging to Cluster 1, Cluster 2, and Cluster 3, respectively. (<b>B</b>) shows the probability density function of Class AB samples clustered into Cluster 1, several high-probability samples with their scalograms (in the upper row), and several low-probability samples with their respective scalograms (in the lower row). (<b>C</b>) shows the probability density functions of Class CD samples clustered into Cluster 3, several high-probability samples with their scalograms, and several low-probability samples with their scalograms (in the lower row).</p>
Full article ">Figure 7
<p>Purity, ARI, and NMI of the results of clustering on four groups of EEG/intracranial EEG (iEEG) data of the Bonn dataset separately using different kinds of similarities.</p>
Full article ">Figure 8
<p>Purity, ARI, and NMI of the results of clustering on three epileptic subjects of ECoG data of the HUP dataset separately using different kinds of similarities.</p>
Full article ">Figure 9
<p>Impact of the iteration number during training W-SLOGAN on the clustering performance on four groups of EEG data of the Bonn dataset separately evaluated with Purity, ARI, and NMI.</p>
Full article ">Figure 10
<p>Impact of the iteration number duringtraining W-SLOGAN on the clustering performance on three epileptic subjects of ECoG data of the HUP dataset separately evaluated with Purity, ARI and NMI.</p>
Full article ">Figure 11
<p>Typical kinds of epileptiform waveforms were found by clustering the ictal iEEG data of the Bonn dataset. In each row are displayed the characteristic waveform of a type of epileptiform discharge, three epileptiform waves of that type that were clustered into a same cluster found from the iEEG recordings by our approach, as well as the baseline scalogram of that cluster.</p>
Full article ">Figure 12
<p>Class labels and clustering results of several samples in group AB_CD_E of the Bonn dataset. Samples on each row belong to a same class and those on each column are clustered into a same cluster. Each grid displays four samples. Row 1 and column 1 both correspond to Class AB, i.e., healthy; Row 2 and column 2 both correspond to Class CD, i.e., inter-ictal, epileptic; Row 3 and column 3 both correspond to Class E, i.e., ictal, epileptic.</p>
Full article ">
19 pages, 7958 KiB  
Article
Extracting Vehicle Trajectories from Partially Overlapping Roadside Radar
by Maxwell Schrader, Alexander Hainen and Joshua Bittle
Sensors 2024, 24(14), 4640; https://doi.org/10.3390/s24144640 - 17 Jul 2024
Cited by 1 | Viewed by 469
Abstract
This work presents a methodology for extracting vehicle trajectories from six partially-overlapping roadside radars through a signalized corridor. The methodology incorporates radar calibration, transformation to the Frenet space, Kalman filtering, short-term prediction, lane-classification, trajectory association, and a covariance intersection-based approach to track fusion. [...] Read more.
This work presents a methodology for extracting vehicle trajectories from six partially-overlapping roadside radars through a signalized corridor. The methodology incorporates radar calibration, transformation to the Frenet space, Kalman filtering, short-term prediction, lane-classification, trajectory association, and a covariance intersection-based approach to track fusion. The resulting dataset contains 79,000 fused radar trajectories over a 26-h period, capturing diverse driving scenarios including signalized intersections, merging behavior, and a wide range of speeds. Compared to popular trajectory datasets such as NGSIM and highD, this dataset offers extended temporal coverage, a large number of vehicles, and varied driving conditions. The filtered leader–follower pairs from the dataset provide a substantial number of trajectories suitable for car-following model calibration. The framework and dataset presented in this work has the potential to be leveraged broadly in the study of advanced traffic management systems, autonomous vehicle decision-making, and traffic research. Full article
(This article belongs to the Special Issue Advances in Intelligent Transportation Systems Based on Sensor Fusion)
Show Figures

Figure 1

Figure 1
<p>Overlay showing three coordinate systems: UTM, Radar Relative, and Frenet.</p>
Full article ">Figure 2
<p>The network of interest, shown with the FoV of all six (6) radars. The colored region is drawn to the 95th percentile of range. The exploded intersection shows the lane center lines used as the Frenet frames. The centroid of the network is located at (33.235, −87.614).</p>
Full article ">Figure 3
<p>Distribution of <span class="html-italic">d</span> measurements from the rightmost lane in both the EB and WB directions, where traffic flow can be considered as coming towards the reader.</p>
Full article ">Figure 4
<p>The process of splitting a vehicle graph into optimal subgraphs. The <b>leftmost</b> graph represents the result of optimistically joining leader–follower pairs; the red edge represents the additional connection that results from analyzing all co-existing tracklets. The <b>middle</b> image displays the result of graph pruning, where the connections between 1 and 3 have been removed along with the connections between 0 and 2 and between 2 and 3, resulting in two separate vehicles. The <b>rightmost</b> subplot displays the <span class="html-italic">s</span>-position of the vehicles.</p>
Full article ">Figure 5
<p>Example of dGPS-equipped probe vehicle matched to radar data in the Frenet frame. The camera image in Subfigure (<b>b</b>) displays one vertical slice of the space–time diagram as indicated in Subfigure (<b>d</b>), which is displaying both of the WB lanes. The leader vehicle is also visible in both the space–time diagram and the image. Subfigure (<b>a</b>) displays the velocity trace of the vehicle as measured by the radar as well as the absolute error resulting from the the CI, Raw, and RTS steps plotted on the secondary y-axis. Subfigure (<b>c</b>) shows the centroid position of the vehicle for all radar processing methods and the corresponding error relative to the dGPS data plotted on the secondary y-axis.</p>
Full article ">Figure 6
<p>Comparison of the RMSE for the front, back, and length of the vehicle, aggregated by distance from the radar. Negative values indicate a vehicle that is approaching the radar, while positive values indicate a vehicle that is moving away from the radar.</p>
Full article ">Figure 7
<p>Time–space diagram showing fused trajectories from EB Lane 1 in <a href="#sensors-24-04640-f002" class="html-fig">Figure 2</a>. Plotted underneath are the light states of the three traffic signals in the network at their corresponding <span class="html-italic">s</span> position. Trajectories that are the same color represent the same vehicle, with colors being recycled every ten vehicles. The black box in the top figure is exploded in the bottom to show more detail.</p>
Full article ">
17 pages, 17666 KiB  
Article
Advanced Integration of Urban Street Greenery and Pedestrian Flow: A Multidimensional Analysis in Chengdu’s Central Urban District
by Qicheng Ma, Jiaxin Zhang and Yunqin Li
ISPRS Int. J. Geo-Inf. 2024, 13(7), 254; https://doi.org/10.3390/ijgi13070254 - 16 Jul 2024
Viewed by 645
Abstract
As urbanization accelerates, urban greenery, particularly street greenery, emerges as a vital strategy for enhancing residents’ quality of life, demanding attention for its alignment with pedestrian flows to foster sustainable urban development and ensure urban dwellers’ wellbeing. The advent of diverse urban data [...] Read more.
As urbanization accelerates, urban greenery, particularly street greenery, emerges as a vital strategy for enhancing residents’ quality of life, demanding attention for its alignment with pedestrian flows to foster sustainable urban development and ensure urban dwellers’ wellbeing. The advent of diverse urban data has significantly advanced this area of study. Focusing on Chengdu’s central urban district, this research assesses street greening metrics against pedestrian flow indicators, employing spatial autocorrelation techniques to investigate the interplay between street greenery and pedestrian flow over time and space. Our findings reveal a prevalent negative spatial autocorrelation between street greenery and pedestrian flow within the area, underscored by temporal disparities in greenery demands across various urban functions during weekdays versus weekends. This study innovatively incorporates mobile phone signal-based population heat maps into the mismatch analysis of street greenery for the first time, moving beyond the conventional static approach of space syntax topology in assessing pedestrian flow. By leveraging dynamic pedestrian flow data, it enriches our understanding of the disconnect between street greening plans and pedestrian circulation, highlighting the concept of urban flow and delving into the intricate nexus among time, space, and human activity. Moreover, this study meticulously examines multiple street usage scenarios, reflecting diverse behavior patterns, with the objective of providing nuanced and actionable strategies for urban renewal initiatives aimed at creating more inviting and sustainable urban habitats. Full article
Show Figures

Figure 1

Figure 1
<p>Research framework for the present study.</p>
Full article ">Figure 2
<p>Overview of the study area: (<b>a</b>) location of Chengdu, (<b>b</b>) location of Chengdu’s central urban district, (<b>c</b>) the elevation of Chengdu’s central urban district.</p>
Full article ">Figure 3
<p>GVI calculating process: (<b>a</b>) an example of street view image acquisition, (<b>b</b>) different calculation results using semantic segmentation.</p>
Full article ">Figure 4
<p>Spatial distribution of research data.</p>
Full article ">Figure 5
<p>NDVI-related bivariate local spatial autocorrelation. The numbers in the legend represent the number of points for each high and low clustering result.</p>
Full article ">Figure 6
<p>GVI-related bivariate local spatial autocorrelation. The numbers in the legend represent the number of points for each high and low clustering result.</p>
Full article ">
16 pages, 1134 KiB  
Article
Developing Problematic Performance Value Scores: Binding Routine Activity Performance, Environmental Barriers, and Health Conditions
by Jimin Choi and JiYoung Park
Int. J. Environ. Res. Public Health 2024, 21(6), 764; https://doi.org/10.3390/ijerph21060764 - 13 Jun 2024
Viewed by 841
Abstract
Background: Community design features, such as sidewalks and street crossings, present significant challenges for individuals with disabilities, hindering their physical performance and social integration. However, limited research has been conducted on the application of Universal Design (UD) to address these challenges, particularly concerning [...] Read more.
Background: Community design features, such as sidewalks and street crossings, present significant challenges for individuals with disabilities, hindering their physical performance and social integration. However, limited research has been conducted on the application of Universal Design (UD) to address these challenges, particularly concerning specific demographic groups and population cohorts. Understanding the influence of environmental features on physical performance is crucial for developing inclusive solutions like UD, which can enhance usability and social integration across diverse populations. Objective: This study aims to bridge this gap by investigating the complex relationships between environmental barriers, health conditions, and routine activity performance. An index was developed to evaluate users’ UD performance based on functional capacity, providing scientifically rigorous and objectively measured evidence of UD effectiveness in creating inclusive built environments. Method: Using data from the Problematic Activities Survey (PAS) conducted in the U.S., Canada, and Australia and targeting individuals with and without functional limitations, multinomial logit models were employed to estimate the probabilities of encountering performance problems. This analysis led to the development of the Problematic Performance Value (PPV) score. Results: The results demonstrated significant disparities in PPVs across various health conditions, particularly concerning curb ramps. Individuals facing mobility issues in their legs/feet, arms/hands, or back/neck encounter more pronounced challenges, especially when curb ramps lack proper design elements. Similarly, individuals with vision impairments face heightened difficulties with traffic signals, particularly due to issues with audible signal systems. These findings underscore the importance of addressing micro-level environmental challenges to accommodate individuals with varying functional capacities effectively. Conclusions: By providing insights into the most problematic daily activities encountered by diverse populations, the PPV score serves as a valuable indicator for guiding environmental design improvements and promoting equitable space usage. This can be used to guide improved UD solutions and decide areas of concentration by providing generalized information on specific environmental features that contribute to user performance. Full article
(This article belongs to the Special Issue Application of Big Data Analysis to Health Risk Assessment)
Show Figures

Figure 1

Figure 1
<p>A conceptual framework for the usability measurement of user performance.</p>
Full article ">Figure 2
<p>Boxplot of PPV scores for different health conditions and environmental barriers (from top, Using Curb Ramps; Using Traffic Signals; and Using Crosswalks).</p>
Full article ">Figure 2 Cont.
<p>Boxplot of PPV scores for different health conditions and environmental barriers (from top, Using Curb Ramps; Using Traffic Signals; and Using Crosswalks).</p>
Full article ">
27 pages, 4669 KiB  
Review
GNSS Reflectometry-Based Ocean Altimetry: State of the Art and Future Trends
by Tianhe Xu, Nazi Wang, Yunqiao He, Yunwei Li, Xinyue Meng, Fan Gao and Ernesto Lopez-Baeza
Remote Sens. 2024, 16(10), 1754; https://doi.org/10.3390/rs16101754 - 15 May 2024
Viewed by 1257
Abstract
For the past 20 years, Global Navigation Satellite System reflectometry (GNSS-R) technology has successfully shown its potential for remote sensing of the Earth’s surface, including ocean and land surfaces. It is a multistatic radar that uses the GNSS signals reflected from the Earth’s [...] Read more.
For the past 20 years, Global Navigation Satellite System reflectometry (GNSS-R) technology has successfully shown its potential for remote sensing of the Earth’s surface, including ocean and land surfaces. It is a multistatic radar that uses the GNSS signals reflected from the Earth’s surface to extract land and ocean characteristics. Because of its numerous advantages such as low cost, multiple signal sources, and all-day/weather and high-spatiotemporal-resolution observations, this new technology has attracted the attention of many researchers. One of its most promising applications is GNSS-R ocean altimetry, which can complement existing techniques such as tide gauging and radar satellite altimetry. Since this technology for ocean altimetry was first proposed in 1993, increasing progress has been made including diverse methods for processing reflected signals (such as GNSS interferometric reflectometry, conventional GNSS-R, and interferometric GNSS-R), different instruments (such as an RHCP antenna with one geodetic receiver, a linearly polarized antenna, and a system of simultaneously used RHCP and LHCP antennas with a dedicated receiver), and different platform applications (such as ground-based, air-borne, or space-borne). The development of multi-mode and multi-frequency GNSS, especially for constructing the Chinese BeiDou Global Navigation Satellite System (BDS-3), has enabled more free signals to be used to further promote GNSS-R applications. The GNSS has evolved from its initial use of GPS L1 and L2 signals to include other GNSS bands and multi-GNSS signals. Using more advanced, multi-frequency, and multi-mode signals will bring new opportunities to develop GNSS-R technology. In this paper, studies of GNSS-R altimetry are reviewed from four perspectives: (1) classifications according to different data processing methods, (2) different platforms, (3) development of different receivers, and (4) our work. We overview the current status of GNSS-R altimetry and describe its fundamental principles, experiments, recent applications to ocean altimetry, and future directions. Full article
(This article belongs to the Special Issue SoOP-Reflectometry or GNSS-Reflectometry: Theory and Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of the GNSS-IR experiment setup.</p>
Full article ">Figure 2
<p><b>Upper</b>: one example of the detrended SNR observations for BDS PRN 08 at B1 frequency at one station. <b>Below</b>: Lomb–Scargle periodogram (LSP) spectral analysis of the detrended SNR observations.</p>
Full article ">Figure 3
<p>Example of two combined multipath errors (<b>top</b> panel) from a dual–frequency code combination for Galileo signals at station AT01 and their correspondent LSP figures (<b>bottom</b> panel), from which we can see two peaks corresponding to dual–frequency signals. The blue and red lines represent two different combined multipath errors.</p>
Full article ">Figure 4
<p>Schematic of (<b>a</b>) ground-based and (<b>b</b>) air-borne GNSS-R altimetry with two antennas.</p>
Full article ">Figure 5
<p>Measured 2-D delay-Doppler maps (<b>a</b>,<b>b</b>) and 1-D delay maps (<b>c</b>,<b>d</b>) for a ground-based (<b>a</b>,<b>c</b>) and a space-borne (<b>b</b>,<b>d</b>) GNSS-R experiment.</p>
Full article ">Figure 6
<p>Geometry of GNSS-R altimetry when (<b>a</b>) considering the Earth as a flat surface, and (<b>b</b>) considering the Earth’s curvature.</p>
Full article ">Figure 7
<p>Schematic of space-borne GNSS-R altimetry with two antennas in which the Earth’s curvature should be considered.</p>
Full article ">Figure 8
<p>Flowchart of the real-time GNSS-R SDR. Modules with different single colors are handled by the GPU and CPU. The mixed-color modules represent data transfer between the CPU and the GPU as depicted in [<a href="#B119-remotesensing-16-01754" class="html-bibr">119</a>].</p>
Full article ">Figure 9
<p>Sea surface heights derived from (<b>a</b>) B1C and (<b>b</b>) B2a code-level measurements with the moving average and radar altimeter data in one coastal GNSS-R experiment as reported in [<a href="#B52-remotesensing-16-01754" class="html-bibr">52</a>].</p>
Full article ">Figure 10
<p>Delay map of the crosstalk signal in one coastal GNSS-R experiment as reported in [<a href="#B120-remotesensing-16-01754" class="html-bibr">120</a>].</p>
Full article ">Figure 11
<p>Reflector heights estimated from the (<b>a</b>) in-harbor and (<b>b</b>) off-shore ship-borne GNSS-R experiments as reported in [<a href="#B51-remotesensing-16-01754" class="html-bibr">51</a>].</p>
Full article ">Figure 12
<p>Equations of the proposed phase combination method for GNSS-R altimetry referenced from [<a href="#B116-remotesensing-16-01754" class="html-bibr">116</a>].</p>
Full article ">Figure 13
<p>Mean revisit times for the four-system GNSS-R simulations at different latitudes during a single cycle at L2 frequencies as found in [<a href="#B63-remotesensing-16-01754" class="html-bibr">63</a>].</p>
Full article ">
18 pages, 3160 KiB  
Article
A Comparative Analysis of the Effect of Orbital Geometry and Signal Frequency on the Ionospheric Scintillations over a Low Latitude Indian Station: First Results from the 25th Solar Cycle
by Ramkumar Vankadara, Nirvikar Dashora, Sampad Kumar Panda and Jyothi Ravi Kiran Kumar Dabbakuti
Remote Sens. 2024, 16(10), 1698; https://doi.org/10.3390/rs16101698 - 10 May 2024
Cited by 1 | Viewed by 1559
Abstract
The equatorial post-sunset ionospheric irregularities induce rapid fluctuations in the phase and amplitude of global navigation satellite system (GNSS) signals which may lead to the loss of lock and can potentially degrade the position accuracy. This study presents a new analysis of L-band [...] Read more.
The equatorial post-sunset ionospheric irregularities induce rapid fluctuations in the phase and amplitude of global navigation satellite system (GNSS) signals which may lead to the loss of lock and can potentially degrade the position accuracy. This study presents a new analysis of L-band scintillation from a low latitude station at Guntur (Geographic 16.44°N, 80.62°E, dip 22.18°), India, for the period of 18 months from August 2021 to January 2023. The observations are categorized either in the medium Earth-orbiting (MEO) or geosynchronous orbiting (GSO) satellites (GSO is considered as a set of the geostationary and inclined geosynchronous satellites) for L1, L2, and L5 signals. The results show a higher occurrence of moderate (0.5 < S4 ≤ 0.8) and strong (S4 > 0.8) scintillations on different signals from the MEO compared to the GSO satellites. Statistically, the average of peak S4 values provides a higher confidence in the severity of scintillations on a given night, which is found to be in-line with the scintillation occurrences. The percentage occurrence of scintillation-affected satellites is found to be higher on L1 compared to other signals, wherein a contrasting higher percentage of affected satellites over GSO than MEO is observed. While a clear demarcation between the L2/L5 signals and L1 is found over the MEO, in the case of GSO, the CCDF over L5 is found to match mostly with the L1 signal. This could possibly originate from the space diversity gain effect known to impact the closely spaced geostationary satellite links. Another major difference of higher slopes and less scatter of S4 values corresponding to L1 versus L2/L5 from the GSO satellite is found compared to mostly non-linear highly scattered relations from the MEO. The distribution of the percentage of scintillation-affected satellites on L1 shows a close match between MEO and GSO in a total number of minutes up to ~60%. However, such a number of minutes corresponding to higher than 60% is found to be larger for GSO. Thus, the results indicate the possibility of homogeneous spatial patterns in a scintillation distribution over a low latitude site, which could originate from the closely spaced GSO links and highlight the role of the number of available satellites with the geometry of the links, being the deciding factors. This helps the ionospheric community to develop inter-GNSS (MEO and GSO) operability models for achieving highly accurate positioning solutions during adverse ionospheric weather conditions. Full article
Show Figures

Figure 1

Figure 1
<p>The IPP trajectory of satellite links for all the available constellations at the study location (KLEF) on 27 August 2022. The location of the GNSS station is marked with a red color dot, whereas the IPP trajectories of MEO, IGSO, and GEO satellites under all constellations are represented by blue, green, and magenta color dots, respectively (see legend).</p>
Full article ">Figure 2
<p>Observed peak S4 per minute is presented row-wise in each columnar panel, which are given in categories of MEO and GSO satellites, respectively, for L1, L2, and L5 for all of the nights. Each peak S4 value is given a unique color in the color bar, corresponding to the occurrence of weak (green, 0.2 &lt; S4 ≤ 0.5), moderate (blue, 0.5 &lt; S4 ≤ 0.8), and strong (red, S4 &gt; 0.8) scintillation.</p>
Full article ">Figure 3
<p>Total S4 occurrences for each night are given in weak (green), moderate (blue), and strong (red) scintillations on the left ordinate axis. Also, the statistically derived average of S4 values from the top-quartile range for each night is shown as an asterisk marker on the right-ordinate axis.</p>
Full article ">Figure 4
<p>The percentage of scintillation-affected satellites per minute from MEO (upper panel) and GSO satellites (lower panel) are shown, respectively, for L1, L2, and L5 in columnar panels. The color bar represents the percentage of satellites affected at a time (ordinate axis) on a night (abscissa).</p>
Full article ">Figure 5
<p>The seasonal percentage S4 distribution is given in terms of intensity (weak-green, moderate-blue, and strong-red,), respectively, for MEO and GSO satellites and their available signals for the seasons during 2021–2023. The seasonal acronyms correspond to SE = September Equinox (August to October); WS = Winter Solstice (November to January); ME = March Equinox (February to April); and SS = Summer Solstice (May to July).</p>
Full article ">Figure 6
<p>Variations in the complementary cumulative distribution function (CCDF) are shown for L1 (blue), L2 (red), and L5 (green) across all the intensities of the S4 binned in 0.01 increments, respectively, for MEO and GSO satellites. Total S4 &gt; 0 occurrence for MEO = 2,788,999 and GSO = 869,275.</p>
Full article ">Figure 7
<p>Comparative distribution of occurrence of S4 &gt; 0.2 on different signals (<b>a</b>,<b>d</b>) L1 and L2 (<b>b</b>,<b>e</b>) L1 and L5 and (<b>c</b>,<b>f</b>) L2 and L5, respectively, given for MEO (upper panels) and GSO (lower panels) satellites. A linear fit is shown by the red line and the respective fit equation is given in each panel in form of <math display="inline"><semantics> <mrow> <mi>Y</mi> <mo>=</mo> <mi>m</mi> <mi>X</mi> <mo>+</mo> <mi>c</mi> </mrow> </semantics></math>, where m represent slope of the fit.</p>
Full article ">Figure 8
<p>Percentage of satellites affected by scintillation (S4 &gt; 0.2) for the number of minutes (log10 scale) is shown for MEO (left panel) and GSO (right panel) L1 with a bin increment of 10%.</p>
Full article ">
10 pages, 2213 KiB  
Article
Interpretable Recurrent Variational State-Space Model for Fault Detection of Complex Systems Based on Multisensory Signals
by Meng Ma and Junjie Zhu
Appl. Sci. 2024, 14(9), 3772; https://doi.org/10.3390/app14093772 - 28 Apr 2024
Viewed by 730
Abstract
It is necessary to develop a health monitoring system (HMS) for complex systems to improve safety and reliability and prevent potential failures. Time-series signals are collected from multiple sensors installed on the equipment that can reflect the health condition of them. In this [...] Read more.
It is necessary to develop a health monitoring system (HMS) for complex systems to improve safety and reliability and prevent potential failures. Time-series signals are collected from multiple sensors installed on the equipment that can reflect the health condition of them. In this study, a novel interpretable recurrent variational state-space model (IRVSSM) is proposed for time-series modeling and anomaly detection. To be specific, the deterministic hidden state of a recursive neural network is used to capture the latent structure of sensor data, while the stochastic latent variables of a nonlinear deep state-space model capture the diversity of sensor data. Temporal dependencies are modeled through a nonlinear transition matrix; an automatic relevance determination network is introduced to selectively emphasize important sensor data. Experimental results demonstrate that the proposed algorithm effectively captures vital information within the sensor data and provides accurate and reliable fault diagnosis during the steady-state phase of liquid rocket engine operation. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of recurrent variational state-space model.</p>
Full article ">Figure 2
<p>Recurrent variational state-space model. (<b>a</b>) Generative model; (<b>b</b>) Inference network.</p>
Full article ">Figure 3
<p>Automatic relevance determination network.</p>
Full article ">Figure 4
<p>(<b>a</b>) Classification accuracy of LRE-1; (<b>b</b>) Classification accuracy of LRE-2.</p>
Full article ">Figure 5
<p>Sensor weight graph.</p>
Full article ">
12 pages, 5657 KiB  
Article
Steering Mirror System with Closed-Loop Feedback for Free-Space Optical Communication Terminals
by Chris Graham, David Bramall, Othman Younus, Amna Riaz, Richard Binns, Eamon Scullion, Robert T. Wicks and Cyril Bourgenot
Aerospace 2024, 11(5), 330; https://doi.org/10.3390/aerospace11050330 - 23 Apr 2024
Cited by 2 | Viewed by 1726
Abstract
Precision beam pointing plays a critical role in free-space optical communications terminals in uplink, downlink and inter-satellite link scenarios. Among the various methods of beam steering, the use of fast steering mirrors (FSM) is widely adopted, with many commercial solutions employing diverse technologies, [...] Read more.
Precision beam pointing plays a critical role in free-space optical communications terminals in uplink, downlink and inter-satellite link scenarios. Among the various methods of beam steering, the use of fast steering mirrors (FSM) is widely adopted, with many commercial solutions employing diverse technologies, particularly focusing on small, high-bandwidth mirrors. This paper introduces a method using lightweight, commercial off-the-shelf components to construct a custom closed-loop steering mirror platform, suitable for mirror apertures exceeding 100 mm. The approach involves integrating optical encoders into two off-the-shelf open-loop actuators. These encoders read the signal reflected on purposefully diamond-machined knurled screw knobs, providing maximum contrast between light and dark lines. The resulting steering mirror has the potential to complement or replace FSM in applications requiring a larger stroke, at the expense of motion speed. In the presented setup, the mirror tilt resolution achieved based on the encoder closed-loop signal feedback is 45 μrad, with a mean slew rate of 1.5 mrad/s. Importantly, the steering assembly is self-locking, requiring no power to maintain a steady pointing angle. Using the mirror to actively correct for a constantly moving incoming beam, a 5-fold increase in concentration of the beam spot on the center of the detector was obtained compared to a fixed position mirror, demonstrating the mirrors ability to correct for satellite platform jitter and drift. Full article
(This article belongs to the Special Issue Space Telescopes & Payloads)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Simplified transceiver terminal using a single optical train, (<b>b</b>) A simplified optical transceiver using a separated optical train design. Red corresponds to the transmitted beam data beam, green to the received data beam, and blue to the received beacon beam.</p>
Full article ">Figure 2
<p>Rendering of a steering mirror assembly, showing a top-down and side view. The assembly uses a bespoke kinematic mount, with a spherical joint enabling motion in the desired axis, controlled by two piezoelectric inertia actuators. The position of the spherical joint is marked in blue, with the positions of the actuators marked in red. The actuators are positioned 82 mm horizontally from the pivot point, and 30 mm vertically on either side of the pivot plane. The mirror has an aperture measuring 130 × 90 mm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Real position of piezo actuator head as a function of piezo step count. (<b>b</b>) Forward drift of 0 step position as the actuator is driven successively between 0 and 40,000 steps.</p>
Full article ">Figure 4
<p>(<b>a</b>) Image of diamond-turned optical encoder track, (<b>b</b>) actuator position plotted against encoder step count for 15 successive movements. The inset shows a magnified view of recorded actuator positions.</p>
Full article ">Figure 5
<p>Rendering of the experimental optical test set-up to measure pointing repeatability and accuracy of mirror assembly.</p>
Full article ">Figure 6
<p>(<b>Left</b>) Repeatability testing of optical encoder feedback on mirror position, with the mirror being driven between 5 points (color coded) in a continuous loop. Arrows show the scanning direction. (<b>Right</b>) Close-up view of one of the rightmost point.</p>
Full article ">Figure 7
<p>Rendering of the experimental optical test set-up for closed-loop control of the steering mirror.</p>
Full article ">Figure 8
<p>(<b>a</b>) Top—uncorrected pitch and yaw variation over 3000 s. Bottom—corrected pitch and yaw with closed loop enabled. (<b>b</b>) Plot of laser spot centroids measured on the camera sensor, with simulated Cubesat platform jitter being provided by a separate fast-steering mirror. The blue plot shows the track of the beam spot on the camera sensor with the piezo-steering mirror switched off, and the orange plot shows the track with real-time corrections switched on.</p>
Full article ">Figure 9
<p>Image of encoder/actuator test set-up inside vacuum chamber.</p>
Full article ">
Back to TopTop