[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

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
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
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
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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (25,685)

Search Parameters:
Keywords = SAR

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 46447 KiB  
Article
Improved Coherent Processing of Synthetic Aperture Radar Data through Speckle Whitening of Single-Look Complex Images
by Luciano Alparone, Alberto Arienzo and Fabrizio Lombardini
Remote Sens. 2024, 16(16), 2955; https://doi.org/10.3390/rs16162955 (registering DOI) - 12 Aug 2024
Abstract
In this study, we investigate the usefulness of the spectral whitening procedure, devised by one of the authors as a preprocessing stage of envelope-detected single-look synthetic aperture radar (SAR) images, in application contexts where phase information is relevant. In the first experiment, each [...] Read more.
In this study, we investigate the usefulness of the spectral whitening procedure, devised by one of the authors as a preprocessing stage of envelope-detected single-look synthetic aperture radar (SAR) images, in application contexts where phase information is relevant. In the first experiment, each of the raw datasets of an interferometric pair of COSMO-SkyMed images, representing industrial buildings amidst vegetated areas, was individually (1) synthesized by the SAR processor without Fourier-domain Hamming windowing; (2) synthesized with Hamming windowing, used to improve the focalization of targets, with the drawback of spatially correlating speckle; and (3) processed for the whitening of complex speckle, using the data obtained in (2). The interferograms were produced in the three cases, and interferometric coherence and phase maps were calculated through 3 × 3 boxcar filtering. In (1), coherence is low on vegetation; the presence of high sidelobes in the system’s point-spread function (PSF) causes the spread of areas featuring high backscattering. In (2), point targets and buildings are better defined, thanks to the sidelobe suppression achieved by the frequency windowing, but the background coherence is abnormally increased because of the spatial correlation introduced by the Hamming window. Case (3) is the most favorable because the whitening operation results in low coherence in vegetation and high coherence in buildings, where the effects of windowing are preserved. An analysis of the phase map reveals that (3) is likely to be facilitated also in terms of unwrapping. Results are presented on a TerraSAR-X/TanDEM-X (TSX-TDX) image pair by processing the interferograms of original and whitened data using a non-local filter. The main results are as follows: (1) with autocorrelated speckle, the estimation error of coherence may attain 16% and inversely depends on the heterogeneity of the scene; and (2) the cleanness and accuracy of the phase are increased by the preliminary whitening stage, as witnessed by the number of residues, reduced by 24%. Benefits are also expected not only for differential InSAR (DInSAR) but also for any coherent analysis and processing carried out performed on SLC data. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of SAR system (onboard sensor and on-ground processor) followed by optional whitening stage.</p>
Full article ">Figure 2
<p>Power spectra: (<b>a</b>) periodogram of SLC data correlated in slant-range direction; (<b>b</b>) frequency response of inverse filter; (<b>c</b>) periodogram of SLC data in (<b>a</b>) after whitening with the filter in (<b>b</b>).</p>
Full article ">Figure 3
<p>Effects of SAR processor: (<b>a</b>) spatially correlated speckle originating from frequency windowing; (<b>b</b>) correlated speckle whitened using the inverse filter in <a href="#remotesensing-16-02955-f002" class="html-fig">Figure 2</a>b; (<b>c</b>) example of a point target focused without a tapering window (negative grayscale).</p>
Full article ">Figure 4
<p>Interferometric pair of images of Pomigliano, master image: (<b>a</b>) processed without a Hamming window; (<b>b</b>) processed with a Hamming window; (<b>c</b>) processed with a Hamming window and subsequently whitened.</p>
Full article ">Figure 5
<p>Coherence maps of Pomigliano estimated using 3 × 3 boxcar filtering: (<b>a</b>) SLC pair processed without a Hamming window; (<b>b</b>) SLC pair processed with a Hamming window; (<b>c</b>) SLC pair processed with a Hamming window and subsequently whitened.</p>
Full article ">Figure 6
<p>Maps of the interferometric phase of Pomigliano estimated using a 3 × 3 sliding window: (<b>a</b>) SLC pair processed without a Hamming window; (<b>b</b>) SLC pair processed with a Hamming window; (<b>c</b>) SLC pair processed with a Hamming window and subsequently whitened.</p>
Full article ">Figure 7
<p>Unfiltered modulus of interferogram of TSX-TDX SLC pair of the Euskirchen test site.</p>
Full article ">Figure 8
<p>Modulus of interferogram of Euskirchen filtered using NL-INSAR: (<b>a</b>) from non-whitened SLC pair; (<b>b</b>) from whitened SLC pair.</p>
Full article ">Figure 9
<p>Coherence maps of Euskirchen estimated using NL-InSAR: (<b>a</b>) from non-whitened SLC pair; (<b>b</b>) from whitened SLC pair.</p>
Full article ">Figure 10
<p>Difference in coherence calculated from whitened and original data of Euskirchen: the overestimation due to correlation reaches 16% in homogeneous areas.</p>
Full article ">Figure 11
<p>Maps of interferometric phases of Euskirchen estimated using NL-InSAR: (<b>a</b>) from non-whitened SLC pair; (<b>b</b>) from whitened SLC pair.</p>
Full article ">Figure 12
<p>Phase residues overlaid on the coherence map of Euskirchen estimated using NL-InSAR: (<b>a</b>) from non-whitened SLC pair; (<b>b</b>) from whitened SLC pair.</p>
Full article ">
4 pages, 1201 KiB  
Correction
Correction: Ziemssen et al. Immune Response to Initial and Booster SARS-CoV-2 mRNA Vaccination in Patients Treated with Siponimod—Final Analysis of a Nonrandomized Controlled Clinical Trial (AMA-VACC). Vaccines 2023, 11, 1374
by Tjalf Ziemssen, Marie Groth, Veronika Eva Winkelmann and Tobias Bopp
Vaccines 2024, 12(8), 911; https://doi.org/10.3390/vaccines12080911 (registering DOI) - 12 Aug 2024
Abstract
The authors would like to make the following corrections to this published paper [...] Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) SARS-CoV-2-specific neutralizing antibody levels in U/mL. (<b>B</b>) SARS-CoV-2-specific serum total antibody levels in U/mL. All the patients with available data were included in the analysis, and individual values are represented by dots. For 11 booster patients, the month 6 visit and the month 1 after the booster visit were identical (cohort 1: n = 7; cohort 2: n = 1; cohort 3: n = 3). The bars show the median values; the black dotted lines indicate assay-specific cut-offs for seropositivity; and the gray dotted lines indicate the maximal value of the quantification range. DMF: dimethyl fumarate; GA: glatiramer acetate; IFN: interferon-beta; n: number of patients with assessments; TF: teriflunomide; and U: units.</p>
Full article ">Figure 2
<p>(<b>A</b>) T-cell response defined as the presence of SARS-CoV-2-reactive T-cells measured by the secretion of either IFN-, IL-2, or both (any level above basal activity); (<b>B</b>) ELISpot-based quantification of T-cell reactivity by calculation of IFN- stimulation indices towards SARS-CoV-2. Each dot represents one patient, and the medians are indicated by horizontal lines. DMF: dimethyl fumarate; GA: glatiramer acetate; IFN: interferon-beta; IFN-: interferon-; n: number of patients with assessments; PBMC: peripheral blood mononuclear cell; and TF: teriflunomide. The T-cell response could not be assessed in three patients with the continued siponimod treatment, one patient in the control group at the month 6 visit, and two patients of cohort 3 at month 1 after the booster because of insufficient cell counts after PBMC isolation. For 11 booster patients, the month 6 visit and the month 1 after the booster visit were identical (cohort 1: n = 7; cohort 2: n = 1; cohort 3: n = 3).</p>
Full article ">
16 pages, 2447 KiB  
Article
Clinical Features and Vaccination Effects among Children with Post-Acute Sequelae of COVID-19 in Taiwan
by Yu-Lung Hsu, Pei-Chi Chen, Yi-Fen Tsai, Chi-Hung Wei, Lawrence Shi-Hsin Wu, Kai-Sheng Hsieh, Miao-Hsi Hsieh, Huan-Cheng Lai, Chien-Heng Lin, Hsiao-Chuan Lin, Chieh-Ho Chen, An-Chyi Chen, Hung-Chih Lin, I-Ching Chou, Wen-Jue Soong, Kao-Pin Hwang, Henry Horng-Shing Lu, Ruby Pawankar, Hui-Ju Tsai and Jiu-Yao Wang
Vaccines 2024, 12(8), 910; https://doi.org/10.3390/vaccines12080910 (registering DOI) - 12 Aug 2024
Abstract
Background: Post-acute sequelae of SARS-CoV-2 infection (PASC) affects patients after recovering from acute coronavirus disease 2019 (COVID-19). This study investigates the impact of SARS-CoV-2 vaccination on PASC symptoms in children in Taiwan during the Omicron pandemic. Methods: We enrolled children under 18 years [...] Read more.
Background: Post-acute sequelae of SARS-CoV-2 infection (PASC) affects patients after recovering from acute coronavirus disease 2019 (COVID-19). This study investigates the impact of SARS-CoV-2 vaccination on PASC symptoms in children in Taiwan during the Omicron pandemic. Methods: We enrolled children under 18 years with PASC symptoms persisting for more than 4 weeks. Data collected included demographics, clinical information, vaccination status, and symptom persistence. We used logistic regression models to compare symptoms in the acute and post-COVID-19 phases and to assess the association between vaccination and these symptoms. Results: Among 500 PASC children, 292 (58.4%) were vaccinated, 282 (52.8%) were male, and the mean (SD) age was 7.6 (4.6) years. Vaccinated individuals exhibited higher odds of experiencing symptoms in the previous acute phase, such as cough (adjusted odds ratio [AOR] = 1.57; 95% confidence interval [CI]: 1.02–2.42), rhinorrhea/nasal congestion (AOR = 1.74; 95% CI: 1.13–2.67), sneezing (AOR = 1.68; 95% CI: 1.02–2.76), sputum production (AOR = 1.91; 95% CI: 1.15–3.19), headache/dizziness (AOR = 1.73; 95% CI: 1.04–2.87), and muscle soreness (AOR = 2.33; 95% CI: 1.13–4.80). In contrast, there were lower odds of experiencing abdominal pain (AOR = 0.49; 95% CI: 0.25–0.94) and diarrhea (AOR = 0.37; 95% CI: 0.17–0.78) in children who had received vaccination during the post-COVID-19 phase. Conclusions: This study revealed clinical features and vaccination effects in PASC children in Taiwan. Vaccination may reduce some gastrointestinal symptoms in the post-COVID-19 phase. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of top ten common clinical symptoms during the acute phase among 500 children with PASC, sorted by age. The heatmap illustrates the frequency of symptoms in children of different ages during the acute phase. Darker red shades represent higher frequencies. The color bar on the right indicates the number of occurrences, ranging from 0 (lightest) to 7 (darkest).</p>
Full article ">Figure 2
<p>Distribution of top ten common clinical symptoms during the phase of the post-COVID-19 condition among 500 children with PASC, sorted by age. The heatmap displays the frequency of symptoms observed in children of different ages during the post-COVID-19 phase. Darker red shades correspond to higher frequencies. The color bar on the right represents the number of occurrences, with values ranging from 0 (lightest) to 5 (darkest).</p>
Full article ">Figure 3
<p><span class="html-italic">P</span>-values distribution of clinical symptoms during the phases of acute and post-COVID-19 conditions for 500 children with PASC. The figure represents the significance of the coincidence between different clinical symptoms during the acute and post-COVID-19 phases. The size and color of the dots represent the <span class="html-italic">p</span>-value: dark blue dots (<span class="html-italic">p</span> &lt; 0.001) indicate a highly significant coincidence, medium blue dots (<span class="html-italic">p</span> = 0.001–0.05) indicate a moderate coincidence, and light blue dots (<span class="html-italic">p</span> ≥ 0.05) indicate a non-significant coincidence.</p>
Full article ">Figure 4
<p>Association between clinical symptoms during the acute phase and vaccination in 500 children with PASC.</p>
Full article ">Figure 5
<p>Association between clinical symptoms during the post-COVID-19 phase and vaccination in 500 children with PASC.</p>
Full article ">
12 pages, 4351 KiB  
Communication
Automatic Estimation of Tropical Cyclone Centers from Wide-Swath Synthetic-Aperture Radar Images of Miniaturized Satellites
by Yan Wang, Haihua Fu, Lizhen Hu, Xupu Geng, Shaoping Shang, Zhigang He, Yanshuang Xie and Guomei Wei
Appl. Sci. 2024, 14(16), 7047; https://doi.org/10.3390/app14167047 (registering DOI) - 11 Aug 2024
Viewed by 336
Abstract
Synthetic-Aperture Radar (SAR) has emerged as an important tool for monitoring tropical cyclones (TCs) due to its high spatial resolution and cloud-penetrating capability. Recent advancements in SAR technology have led to smaller and lighter satellites, yet few studies have evaluated their effectiveness in [...] Read more.
Synthetic-Aperture Radar (SAR) has emerged as an important tool for monitoring tropical cyclones (TCs) due to its high spatial resolution and cloud-penetrating capability. Recent advancements in SAR technology have led to smaller and lighter satellites, yet few studies have evaluated their effectiveness in TC monitoring. This paper employs an algorithm for automatic TC center location, involving three stages: coarse estimation from a whole SAR image; precise estimation from a sub-SAR image; and final identification of the center using the lowest Normalized Radar Cross-Section (NRCS) value within a smaller sub-SAR image. Using three wide-swath miniaturized SAR images of TC Noru (2022), and TCs Doksuri and Koinu (2023), the algorithm’s accuracy was validated by comparing estimated TC center positions with visually located data. For TC Noru, the distances for the three stages were 21.42 km, 14.39 km, and 8.19 km; for TC Doksuri—14.36 km, 20.48 km, and 17.10 km; and for TC Koinu—47.82 km, 31.59 km, and 5.42 km. The results demonstrate the potential of miniaturized SAR in TC monitoring. Full article
(This article belongs to the Topic Radar Signal and Data Processing with Applications)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the TC-center estimation algorithm. <b><span class="html-italic">β</span></b> represents the compensation angle. The coarsely estimated center position is derived from Stage 1, the precisely estimated center position is derived from Stage 2, and NRCS-adjusted center position is derived from the optional stage.</p>
Full article ">Figure 2
<p>The rule for determining whether a line passes through a point. The green points represent a SSWD point. The green solid line depicts the retrieved SSWD, and the green dashed line represents the compensated SSWD, with <span class="html-italic">β</span> indicating the compensation angle. The black points represent candidate points, and the interval is <span class="html-italic">M</span><sub>1</sub> equal to the diameter of the circles. If the distance from a candidate point to a line is less than M1/2, the line is considered to pass through the candidate point. The red circles indicate that the line passes through their centered candidate points.</p>
Full article ">Figure 3
<p>Variation in maximum values in heatmaps by compensation angles. The vertical blue and orange lines represent the compensation angle corresponding to the peak values of the curves. (<b>a</b>) TC Noru; (<b>b</b>) TC Doksuri; (<b>c</b>) TC Koinu. The global maximum values of the first stage (Stage 1) occurred at 0.5°, −16.5°, and −17.5°, respectively, and the corresponding heat maps are shown in <a href="#applsci-14-07047-f004" class="html-fig">Figure 4</a>; the global maximum values of the second stage (Stage 2) occurred at −0.5°, −27.5°, and −23.0°, respectively, and the corresponding heat maps are shown in <a href="#applsci-14-07047-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 4
<p>Coarse estimation for centers of (<b>a</b>,<b>d</b>) TC Noru; (<b>b</b>,<b>e</b>) TC Doksuri; (<b>c</b>,<b>f</b>) TC Koinu. The coarsely estimated TC center positions are denoted by red points, while the visually located TC center positions are denoted by cyan triangles in (<b>a</b>–<b>c</b>). The sub-SAR images inside the green boxes are used to precisely estimate the TC center position. (<b>d</b>–<b>f</b>) represent the heatmaps corresponding to the compensation angles of 0.5°, −16.5°, and −17.5°, respectively. The position of the maximum value in (<b>d</b>–<b>f</b>) represents the coarsely estimated TC center positions (red points in (<b>a</b>–<b>c</b>)).</p>
Full article ">Figure 5
<p>Precise estimation for centers of (<b>a</b>,<b>d</b>) TC Noru; (<b>b</b>,<b>d</b>) TC Doksuri; (<b>c</b>,<b>f</b>) TC Koinu. The precisely estimated TC center positions are denoted by green points in (<b>a</b>–<b>c</b>). The sub-SAR images inside the blue boxes are used to adjust the precise estimation by NRCS. (<b>d</b>–<b>f</b>) represent the heatmaps corresponding to the compensation angles of −0.5°, −27.5°, and −23.0°, respectively. The position of the maximum value in (<b>d</b>–<b>f</b>) represents the precisely estimated TC center positions (green points in (<b>a</b>–<b>c</b>)).</p>
Full article ">Figure 6
<p>NRCS-adjusted estimation for centers of (<b>a</b>) TC Noru; (<b>b</b>) TC Doksuri; (<b>c</b>) TC Koinu. The NRCS-adjusted TC centers are denoted by blue points, with the point corresponding to the lowest NRCS value.</p>
Full article ">
35 pages, 11986 KiB  
Article
Modeling Ocean Swell and Overtopping Waves: Understanding Wave Shoaling with Varying Seafloor Topographies
by Chak-Nang Wong and Kwok-Wing Chow
J. Mar. Sci. Eng. 2024, 12(8), 1368; https://doi.org/10.3390/jmse12081368 - 11 Aug 2024
Viewed by 265
Abstract
One risk posed by hurricanes and typhoons is local inundation as ocean swell and storm surge bring a tremendous amount of energy and water flux to the shore. Numerical wave tanks are developed to understand the dynamics computationally. The three-dimensional equations of motion [...] Read more.
One risk posed by hurricanes and typhoons is local inundation as ocean swell and storm surge bring a tremendous amount of energy and water flux to the shore. Numerical wave tanks are developed to understand the dynamics computationally. The three-dimensional equations of motion are solved by the software ‘Open Field Operation And Manipulation’ v2206. The ‘Large Eddy Simulation’ scheme is adopted as the turbulence model. A fifth-order Stokes wave is taken as the inlet condition. Breaking, ‘run-up’, and overtopping waves are studied for concave, convex, and straight-line seafloors for a fixed ocean depth. For small angles of inclination (<10°), a convex seafloor displays wave breaking sooner than a straight-line one and thus actually delivers a smaller volume flux to the shore. Physically, a convex floor exhibits a greater rate of depth reduction (on first encounter with the sloping seafloor) than a straight-line one. Long waves with a speed proportional to the square root of the depth thus experience a larger deceleration. Nonlinear (or ‘piling up’) effects occur earlier than in the straight-line case. All these scenarios and reasoning are reversed for a concave seafloor. For large angles of inclination (>30°), impingement, reflection, and deflection are the relevant processes. Empirical dependence for the setup and swash values for a convex seafloor is established. The reflection coefficient for waves reflected from the seafloor is explored through Fourier analysis, and a set of empirical formulas is developed for various seafloor topographies. Understanding these dynamical factors will help facilitate the more efficient designing and construction of coastal defense mechanisms against severe weather. Full article
(This article belongs to the Special Issue Hydrodynamic Research of Marine Structures)
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the numerical wave tank.</p>
Full article ">Figure 2
<p>(<b>a</b>) Location of Heng Fa Chuen as indicated by the cross, and (<b>b</b>) bathymetric chart of the Heng Fa Chuen neighborhood in Hong Kong from Hong Kong Hydrographic Office.</p>
Full article ">Figure 3
<p>The computational domain is divided into two regions, namely (A) wave generation and relaxation zone and (B) region of computation investigation. (<b>a</b>) Mesh design in region A. (<b>b</b>) Mesh design in region B.</p>
Full article ">Figure 3 Cont.
<p>The computational domain is divided into two regions, namely (A) wave generation and relaxation zone and (B) region of computation investigation. (<b>a</b>) Mesh design in region A. (<b>b</b>) Mesh design in region B.</p>
Full article ">Figure 4
<p>Computational domain and nomenclature of the boundary patches.</p>
Full article ">Figure 5
<p>Schematic diagram of the concave, straight-line, and convex seafloors.</p>
Full article ">Figure 6
<p>Surface elevations recorded at Probe 1 for three different spatial discretization schemes.</p>
Full article ">Figure 7
<p>Surface elevations recorded at Probe 1 with different Courant numbers (‘CourN’ in the inset).</p>
Full article ">Figure 8
<p>Normalized volume of water per unit channel width of the (<b>a</b>) straight-line seafloor, (<b>b</b>) wave setup, and (<b>c</b>) swash of different seafloor geometries at various angles.</p>
Full article ">Figure 9
<p>(<b>a</b>) Averaged normalized volume fluxes per unit width of the channel at the outlet and (<b>b</b>) averaged normalized surface elevations of different seafloors with various angles of the slopes.</p>
Full article ">Figure 10
<p>Wave breaking process on a straight-line seafloor at a slope inclination of 4.6° at 58° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (<b>d</b>) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 11
<p>Wave breaking process on a straight-line seafloor at a slope inclination of 8.1° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (<b>d</b>) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 12
<p>Wave breaking process on a concave seafloor at a slope inclination of 4.6° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (<b>d</b>) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 13
<p>Wave breaking process on a concave seafloor at a slope inclination of 8.1° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (<b>d</b>) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 14
<p>Nonlinearities at different positions (<span class="html-italic">x*</span>) on various seafloor geometries at 4.6° inclination.</p>
Full article ">Figure 15
<p>Wave breaking process on a convex seafloor at a slope inclination of 4.6° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (<b>d</b>) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 16
<p>Wave breaking process on a convex seafloor at a slope inclination of 8.1° at the normalized time instants of (<b>a</b>) 109.6, (<b>b</b>) 111.2, (<b>c</b>) 112.8, (d) 114.3 (<b>e</b>) 115.9, and (<b>f</b>) 117.5.</p>
Full article ">Figure 17
<p>Instantaneous normalized velocity magnitudes of (<b>a</b>) straight-line seafloor at 4.6°, (<b>b</b>) concave seafloor at 4.6°, (<b>c</b>) convex seafloor at 4.6°, (<b>d</b>) straight-line seafloor at 58°, (<b>e</b>) concave seafloor at 58°, and (<b>f</b>) convex seafloor at 58° at the normalized time instant of 117.5.</p>
Full article ">Figure 18
<p>Normalized turbulent kinetic energy fields of (<b>a</b>) straight-line seafloor at 4.6°, (<b>b</b>) concave seafloor at 4.6°, (<b>c</b>) convex seafloor at 4.6°, (<b>d</b>) straight-line seafloor at 58°, (<b>e</b>) concave seafloor at 58°, and (<b>f</b>) convex seafloor at 58° at the normalized time instant of 117.5.</p>
Full article ">Figure 19
<p>Normalized surface elevation of (<b>a</b>) straight-line seafloor at 4.6°, (<b>b</b>) concave seafloor at 4.6°, (<b>c</b>) convex seafloor at 4.6°, (<b>d</b>) straight-line seafloor at 58°, (<b>e</b>) concave seafloor at 58°, and (<b>f</b>) convex seafloor at 58° at the normalized time instant of 117.5.</p>
Full article ">Figure 20
<p>The fitted curve of (<b>a</b>) the setup and (<b>b</b>) the swash of the convex seafloor.</p>
Full article ">Figure 21
<p>The reflection coefficient of different seafloors with various Iribarren numbers.</p>
Full article ">Figure 22
<p>Schematic diagram of wave propagation on very steep (<b>a</b>) straight-line, (<b>b</b>) concave, and (<b>c</b>) convex seafloors.</p>
Full article ">Figure 23
<p>(<b>a</b>) The schematic diagram of the effective slope for a convex seafloor and (<b>b</b>) the normalized volume of water per unit channel width across the outlet patch of the convex seafloor and the hypothetical slope.</p>
Full article ">
13 pages, 5142 KiB  
Article
Spike Protein of SARS-CoV-2 Activates Cardiac Fibrogenesis through NLRP3 Inflammasomes and NF-κB Signaling
by Huynh Van Tin, Lekha Rethi, Satoshi Higa, Yu-Hsun Kao and Yi-Jen Chen
Cells 2024, 13(16), 1331; https://doi.org/10.3390/cells13161331 - 11 Aug 2024
Viewed by 734
Abstract
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. [...] Read more.
Background: The spike protein of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) is crucial to viral entry and can cause cardiac injuries. Toll-like receptor 4 (TLR4) and NOD-, LPR-, and pyrin-domain-containing 3 (NLRP3) inflammasome are critical immune system components implicated in cardiac fibrosis. The spike proteins activate NLRP3 inflammasome through TLR4 or angiotensin-converting enzyme 2 (ACE2) receptors, damaging various organs. However, the role of spike proteins in cardiac fibrosis in humans and the interactions of spike proteins with NLRP3 inflammasomes and TLR4 remain poorly understood. Methods: We utilized scratch assays, Western blotting, and immunofluorescence to evaluate the migration, fibrosis signaling, mitochondrial calcium levels, reactive oxygen species (ROS) production, and cell morphology of cultured human cardiac fibroblasts (CFs) treated with spike (S1) proteins for 24 h with or without an anti-ACE2 neutralizing antibody, a TLR4 blocker, or an NLRP3 inhibitor. Results: S1 protein enhanced CFs migration and the expressions of collagen 1, α-smooth muscle actin, transforming growth factor β1 (TGF-β1), phosphorylated SMAD2/3, interleukin 1β (IL-1β), and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB). S1 increased ROS production but did not affect mitochondrial calcium content and cell morphology. Treatment with an anti-ACE2 neutralizing antibody attenuated the effects of S1 on collagen 1 and TGF-β1 expressions. Moreover, NLRP3 (MCC950) and NF-kB inhibitors, but not the TLR4 inhibitor TAK-242, prevented the S1-enhanced CFs migration and overexpression of collagen 1, TGF-β1, and IL-1β. Conclusion: S1 activates human CFs by priming NLRP3 inflammasomes through NF-κB signaling in an ACE2-dependent manner. Full article
(This article belongs to the Special Issue Insight into Cardiomyopathy)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>S1 protein enhanced CFs activation. Treatment with S1 protein (5 nM) for 24 h increased cell migration (<b>A</b>) but not cell proliferation (<b>B</b>) in CFs. (<b>C</b>) Additionally, S1 protein also elevated pro-COL1A1 and α-SMA protein expressions. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 2
<p>S1 protein increased profibrotic signaling. Treatment with S1 protein (5 nM) for 24 h increased protein expressions of TGF-β1 and pSMAD2/3 measured using Immunoblot ((<b>A</b>) <span class="html-italic">n</span> = 4 independent experiments), secretion of TGF-β1 in cultured medium measured using ELISA assays ((<b>B</b>) <span class="html-italic">n</span> = 5 independent experiments), and TGF-β1 mRNA quantified using RT-qPCR in CFs ((<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 3
<p>Impact of NLRP3 and TLR4 signaling on S1 protein-mediated CFs migration. Pretreatment with MCC950 (10 µM (<b>A</b>)) but not TAK-242 (1 µM (<b>B</b>)) blocked the effects of S1 protein (5 nM) treatment for 24 h on cell migration. <span class="html-italic">n</span> = 3 independent experiments.</p>
Full article ">Figure 4
<p>The role of NLRP3 signaling on S1 protein-induced fibrotic markers in CFs. MCC950 effectively blocked the effect of S1 on expressions of fibrotic markers including pro-COL1A1, TGF-β1, and its downstream target, pSMAD2/3. <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 5
<p>Role of TLR4 signaling on S1 protein-induced fibrotic markers in CFs. TAK-242 did not change the effect of S1 on pro-COL1A1 and TGF-β1 expressions in CFs (10 µM, B). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 6
<p>Role of NF-κB on S1 protein-mediated CFs migration and expressions of fibrotic factors. Pretreatment with BAY 11-7082 (an NF-κB inhibitor, 3 µM) completely blocked the effects of S1 protein (5 nM for 24 h) on CFs migration ((<b>A</b>) <span class="html-italic">n</span> = 5 independent experiments) and the protein expressions of IL1-β cleavage, TGF-β1, and pSMAD2/3 ((<b>B</b>,<b>C</b>) <span class="html-italic">n</span> = 4 independent experiments).</p>
Full article ">Figure 7
<p>S1 activated CFs in an ACE2-dependent manner. ACE2 neutralization antibody (5 µM) effectively blocked the enhanced CFs migration induced by the S1 protein (5 nM for 24 h) (<b>A</b>), along with suppressing the protein expressions of pro-COL1A1, TGF-β1, and phosphorylated NF-κB (p-p65) (<b>B</b>). <span class="html-italic">n</span> = 4 independent experiments.</p>
Full article ">Figure 8
<p>Mechanisms underlying S1 protein-induced activation of CFs and promotion of ECM protein synthesis. Key processes involve activation of the NLRP3 inflammasome, ACE2/NF-κB signaling, and ROS formation. Upon binding to ACE2, the S1 protein initiates a signaling cascade that activates NF-κB, a transcription factor promoting the expression of inflammation-related genes, including those required for NLRP3 inflammasome and pro-interleukin (IL)-1β. ROS formation triggers NLRP3 inflammasome activation, leading to processing of pro-IL-1β into its mature form by caspase-1. Mature IL-1β is released extracellularly, binds to its receptor, and initiates a signaling cascade, enhancing TGF-β1 production, promoting CFs activation, and inducing ECM synthesis, ultimately contributing to cardiac fibrosis. Abbreviations: SARS-CoV-2: severe acute respiratory syndrome coronary virus 2, CFs: cardiac fibroblasts, ECM: extra cellular matrix, EMT: epithelial-mesenchymal transition, NLRP3: NLR family pyrin domain containing 3, ACE2: angiotensin converting enzyme 2, NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells, ROS: reactive oxygen species, IL-1β: Interleukin 1 beta.</p>
Full article ">
12 pages, 1106 KiB  
Article
Lung Ultrasound Efficacy in Monitoring Post-SARS-CoV-2 Pneumonia and Inflammatory Biomarkers in Pediatric Patients
by Ramona Chelcea, Mihaela Dediu, Diana Dabica, Sorina Maria Denisa Laitin and Ioana Mihaiela Ciuca
Medicina 2024, 60(8), 1296; https://doi.org/10.3390/medicina60081296 - 11 Aug 2024
Viewed by 382
Abstract
Background and Objectives: Recognizing the crucial gaps in our understanding of pediatric pneumonia post-SARS-CoV-2 infection, this study aimed to assess the relationship between Pediatric Pneumonia Ultrasound Scores (PedPne) and inflammatory biomarkers. The primary objective of this study is to evaluate the predictive [...] Read more.
Background and Objectives: Recognizing the crucial gaps in our understanding of pediatric pneumonia post-SARS-CoV-2 infection, this study aimed to assess the relationship between Pediatric Pneumonia Ultrasound Scores (PedPne) and inflammatory biomarkers. The primary objective of this study is to evaluate the predictive value of PedPne in comparison with inflammatory biomarkers (IL-6 and dNLR) for the development of pneumonia in pediatric patients following SARS-CoV-2 infection. Materials and Methods: This longitudinal observational study collected data from pediatric patients diagnosed with pneumonia after an acute SARS-CoV2 infection. The study focused on analyzing changes in PedPne scores and inflammatory markers such as IL-6 and dNLR from initial admission to follow-up at 7 days. Statistical analysis involved calculating the sensitivity, specificity, and Area Under the Curve (AUC) for each biomarker, alongside regression analysis to determine their hazard ratios for predicting pneumonia development. Results: The analysis identified significant cutoff values for dNLR at 1.88 (sensitivity 77.0%, specificity 85.7%, AUC 0.802, p < 0.001), IL-6 at 6.1 pg/mL (sensitivity 70.3%, specificity 92.9%, AUC 0.869, p < 0.001), and PedPne score at 3.3 (sensitivity 75.7%, specificity 78.6%, AUC 0.794, p < 0.001). Conversely, NLR showed lower diagnostic performance (AUC 0.485, p = 0.327). Regression analysis further highlighted the strong predictive power of these markers, with IL-6 showing a fourfold increase in pneumonia risk (HR = 4.25, CI: 2.07–9.53, p < 0.001), dNLR indicating more than a twofold increase (HR = 2.53, CI: 1.19–6.97, p = 0.006), and PedPne score associated with more than a doubling of the risk (HR = 2.60, CI: 1.33–5.18, p < 0.001). Conclusions: The study conclusively demonstrated that both PedPne ultrasound scores and specific inflammatory biomarkers such as dNLR and IL-6 are significant predictors of pneumonia development in pediatric patients post-COVID-19 infection. These findings advocate for the integration of these biomarkers in routine clinical assessments to enhance the diagnostic accuracy and management of pneumonia in children following SARS-CoV-2 infection. Full article
Show Figures

Figure 1

Figure 1
<p>ROC analysis for pneumonia development among children admitted with COVID-19.</p>
Full article ">Figure 2
<p>Regression analysis for pneumonia development among children admitted for COVID-19.</p>
Full article ">
11 pages, 1127 KiB  
Article
Introducing a Simple Tool of Patient Self-Assessment of Wrist Range of Motion
by Maximilian C. Stumpfe, Kaya Beneke, Raymund E. Horch, Andreas Arkudas, Wibke Müller-Seubert and Aijia Cai
Life 2024, 14(8), 997; https://doi.org/10.3390/life14080997 (registering DOI) - 10 Aug 2024
Viewed by 260
Abstract
Hand disorders can reduce wrist range of motion (ROM). The SARS-CoV-2 pandemic highlighted challenges in routine follow-up exams, making telemedicine a viable solution. This study evaluates the feasibility and accuracy of patient self-measured wrist ROM using a self-designed goniometer template. The template was [...] Read more.
Hand disorders can reduce wrist range of motion (ROM). The SARS-CoV-2 pandemic highlighted challenges in routine follow-up exams, making telemedicine a viable solution. This study evaluates the feasibility and accuracy of patient self-measured wrist ROM using a self-designed goniometer template. The template was designed to measure flexion/extension and radial/ulnar abduction movements. A cohort of 50 adults (25 males/25 females) participated in this prospective study. The exclusion criteria included wrist immobilization and ages outside of 18–65 years. Participants self-assessed their wrist ROM with the goniometer template. Measurements were independently performed by a student and a specialist using standard goniometry, as well as a resident using the self-designed goniometer. The results were blinded for unbiased analysis. Mean differences in ROM varied across movement directions, with minimal differences for ulnar abduction and more substantial deviations for radial abduction, extension and flexion. The patient–specialist comparison showed deviations below 5 degrees for flexion and ulnar abduction in 50% of cases. Telemedicine, expanded by the COVID-19 pandemic, offers significant potential for hand rehabilitation. Current methods of ROM assessment lack cost-effectiveness and simplicity. Our method, demonstrating comparable accuracy for most movements, provides a cost-effective, reliable alternative for remote ROM assessment, enhancing telemedicine practices in hand rehabilitation. Full article
Show Figures

Figure 1

Figure 1
<p>Self-designed goniometer template to determine the range of motion.</p>
Full article ">Figure 2
<p>Overview of the assessment methods and the designated persons responsible for the measurements (created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 28 June 2024).</p>
Full article ">Figure 3
<p>Graphical illustration of the ROM, divided into the individual examiners, movement and sides.</p>
Full article ">
16 pages, 69780 KiB  
Article
The 2024 Mw 7.1 Wushi Earthquake: A Thrust and Strike-Slip Event Unveiling the Seismic Mechanisms of the South Tian Shan’s Thick-Skin Tectonics
by Jiangtao Qiu, Jianbao Sun and Lingyun Ji
Remote Sens. 2024, 16(16), 2937; https://doi.org/10.3390/rs16162937 (registering DOI) - 10 Aug 2024
Viewed by 396
Abstract
The southern margin of the South Tian Shan has drawn attention due to the intense compressional deformation and seismic activity associated with its thrust structures. However, the deformation and seismic activity in the thick-skinned thrust sheets of the root zones are minimal. The [...] Read more.
The southern margin of the South Tian Shan has drawn attention due to the intense compressional deformation and seismic activity associated with its thrust structures. However, the deformation and seismic activity in the thick-skinned thrust sheets of the root zones are minimal. The Mw 7.1 Wushi earthquake on 23 January 2024 serves as a window to reveal these unknown aspects of the seismic mechanisms in this structural setting. Using the Leveraging Interferometric Synthetic Aperture Radar (InSAR) technique, we unlock critical insights into the coseismic deformation fields. The seismogenic fault is an unmapped segment within the Maidan Fault system, exhibiting a strike ranging from 241° to 222°. It is characterized by a shallow dip angle of 62° and a deeper dip angle of 56°. Remarkably, the seismic rupture did not propagate to the Earth’s surface. The majority of slip distribution is concentrated within a range of 4 to 26 km along the strike, indicating that this earthquake was a thrust event on a blind fault within the thick-skinned tectonics of the South Tian Shan. Coulomb stress changes indicate that aftershocks primarily occur in the stress-loading region. Interestingly, some aftershocks are very shallow, causing clear surface deformation. Inversion results show that the fault planes of two aftershocks are located above the main shock fault plane at extremely shallow depths (<6 km). Combining geophysical profile data, we infer that ruptures in the deep-seated thick-skinned structures during the main shock triggered ruptures in the shallow thrust structures. This triggering relationship highlights the potential for combined ruptures of the main shocks and aftershocks in the deep-seated thick-skinned structures beneath the South Tian Shan to result in larger disasters than typical seismic events. Full article
(This article belongs to the Special Issue Monitoring Geohazard from Synthetic Aperture Radar Interferometry)
Show Figures

Figure 1

Figure 1
<p>Major Cenozoic faults in the Tianshan Zone and the tectonic setting of the 2024 Wushi earthquake. (<b>a</b>) The blue arrow indicates the GPS horizontal velocity field from [<a href="#B9-remotesensing-16-02937" class="html-bibr">9</a>]; the red arrow indicates the GPS horizontal velocity field from [<a href="#B10-remotesensing-16-02937" class="html-bibr">10</a>]; the thin black lines represent faults; the empty circles represent the locations of earthquakes with a magnitude of 6 or higher since 1900. (<b>b</b>) The red circular region delineates the area shown in (<b>a</b>); the gray circles represent the locations of strong earthquakes with a magnitude of 7 or higher since 1900 (data are from <a href="https://earthquake.usgs.gov/earthquakes/" target="_blank">https://earthquake.usgs.gov/earthquakes/</a>, accessed on 23 February 2024). (<b>c</b>) shows different organizations’ determined focal mechanism solutions and locations. (<b>d</b>) North–south structural diagram cross-section of the southern margin of the South Tian Shan. The light-blue line delineates the detachment fault. The red lines indicate the thrust faults of the southern margin of the South Tian Shan. Modified from [<a href="#B11-remotesensing-16-02937" class="html-bibr">11</a>,<a href="#B12-remotesensing-16-02937" class="html-bibr">12</a>]. The lemon chiffon area delineates thin skinned structure, medium purple area delineates thick skinned structure. Fault abbreviation: SNBF, the South Naryn Basin Fault; NNBF, the North Naryn Basin Fault; SIKF, the South Issyk-Kul Fault; PFF, the Pamir Frontal Thrust Fault; TFF, Talas-Fergana Fault; MDF, Maidan Fault; TSF, Toshgan Fault; KKSF, Kokesale Fault; NWSF the North Wensu Fault; KTF, Kepingtag Fault.</p>
Full article ">Figure 2
<p>Unwrapped InSAR deformation fields of the Wushi earthquake and aftershocks. (<b>a</b>,<b>b</b>) LOS deformation (cold colors indicate motion away from the satellite, while warm colors indicate motion towards the satellite). White lines denote of the surface trace of the seismogenic fault that we inferred. (<b>c</b>,<b>d</b>) Deformation profiles along A-A’ and B-B’ in (<b>a</b>,<b>b</b>). (<b>e</b>,<b>f</b>) InSAR deformation fields of aftershocks. The blue circles represent precise locations of aftershocks with depths less than 10 km and magnitudes greater than M4.5 between 24 January 2024 and 7 February 2024. The size of the circles corresponds to the magnitude of the aftershocks. Data source: <a href="https://data.earthquake.cn/gxdt/info/2024/334671642.html" target="_blank">https://data.earthquake.cn/gxdt/info/2024/334671642.html</a>, (accessed on 12 March 2024). (<b>e</b>) Ascending track 56, time interval 20240125_20240207. (<b>f</b>) Descending track 34, time interval 20240124_20240206.</p>
Full article ">Figure 3
<p>Marginal posterior probability distributions for the fault model parameters for the Wushi earthquake. Red lines represent the maximum a posteriori probability solution (cold colors for low frequency, warm colors for high frequency).</p>
Full article ">Figure 4
<p>Slip distribution model and deformation and residuals predicted of the Wushi Mw7.1 earthquake. (<b>a</b>) 3D display and (<b>b</b>) 2D display, arrows indicate the slip direction of the hanging wall relative to the footwall. (<b>c</b>,<b>f</b>) represent the observed values of ascending track 56 and descending track 34 after downsampling; (<b>d</b>,<b>g</b>) represent the predicted values; (<b>e</b>,<b>h</b>) represent the residual values. The red solid line indicates the determined trace of seismogenic fault.</p>
Full article ">Figure 5
<p>Estimated slip distribution and predicted deformation of the aftershocks. (<b>a</b>) A three-dimensional visualization showcasing the slip distribution of two faults. Red circles depict aftershock events occurring during the SAR imagery acquisition period. (<b>b</b>,<b>c</b>) Two-dimensional representations of the same slip distribution for enhanced clarity. (<b>d</b>–<b>f</b>) Observed deformation, simulated deformation, and residual errors derived from ascending track T56, respectively. (<b>g</b>–<b>i</b>) Similarly, the observed deformation, simulated deformation, and residual errors for descending track T34 are presented in sequence. The thin blue lines represent the seismogenic fault of the Wushi Mw 7.1 earthquake. The red solid line indicates the surface trace of f1 and f2 faults.</p>
Full article ">Figure 6
<p>(<b>a</b>) Coulomb stress triggering and (<b>b</b>) source structure locations of the main shock–aftershock sequence of the Wushi earthquake. (<b>a</b>) Changes in positive Coulomb stress (depicted in red) on the fault plane indicate proximity to failure and sliding hazard, while negative values (depicted in blue) signify a lack of sliding hazard. (<b>b</b>) The black lines represent the fault locations delineated based on the inversion of fault geometry parameters and geological cross-section base map (adapted from [<a href="#B27-remotesensing-16-02937" class="html-bibr">27</a>]). The red lines depict faults, the black line segments represent the mainshock fault and the aftershock (f1) fault identified in this study. Shallow light yellow areas indicate reverse-thrust overlying strata.</p>
Full article ">
14 pages, 2179 KiB  
Article
Genetic, Clinical, Epidemiological, and Immunological Profiling of IgG Response Duration after SARS-CoV-2 Infection
by Flávia Póvoa da Costa, Kevin Matheus Lima de Sarges, Rosilene da Silva, Erika Ferreira dos Santos, Matheus Holanda do Nascimento, Alice Maciel Rodrigues, Marcos Henrique Damasceno Cantanhede, Fabíola Brasil Barbosa Rodrigues, Maria de Nazaré do Socorro de Almeida Viana, Mauro de Meira Leite, Camille Ferreira de Oliveira, Pablo Fabiano Moura das Neves, Gabriel dos Santos Pereira Neto, Mioni Thieli Figueiredo Magalhães de Brito, Andréa Luciana Soares da Silva, Daniele Freitas Henriques, Juarez Antônio Simões Quaresma, Luiz Fábio Magno Falcão, Maria Alice Freitas Queiroz, Izaura Maria Vieira Cayres Vallinoto, Antonio Carlos Rosário Vallinoto, Giselle Maria Rachid Viana and Eduardo José Melo dos Santosadd Show full author list remove Hide full author list
Int. J. Mol. Sci. 2024, 25(16), 8740; https://doi.org/10.3390/ijms25168740 (registering DOI) - 10 Aug 2024
Viewed by 431
Abstract
The IgG response against SARS-CoV-2 infection can persist for over six months (long response; LR). However, among 30% of those infected, the duration can be as short as three months or less (short response; SR). The present study assembled serological data on the [...] Read more.
The IgG response against SARS-CoV-2 infection can persist for over six months (long response; LR). However, among 30% of those infected, the duration can be as short as three months or less (short response; SR). The present study assembled serological data on the anti-SARS-CoV-2 IgG response duration of two previous studies and integrated these results with the plasmatic cytokine levels and genetic profile of 10 immune-relevant SNPs that were also previously published, along with the plasmatic total IgG, IgA, and IgM levels, allowing for the genetic, clinical, immunological, and epidemiological aspects of the post-COVID-19 IgG response duration to be understood. The SR was associated with previous mild acute COVID-19 and with an SNP (rs2228145) in IL6R related to low gene expression. Additionally, among the SR subgroup, no statistically significant Spearman correlations were observed between the plasma levels of IL-17A and the Th17 regulatory cytokines IFN-γ (rs = 0.2399; p = 0.1043), IL-4 (rs = 0.0273; p = 0.8554), and IL-2 (rs = 0.2204; p = 0.1365), while among the LR subgroup, weaker but statistically significant Spearman correlations were observed between the plasma levels of IL-17A and IFN-γ (rs = 0.3873; p = 0.0016), IL-4 (rs = 0.2671; p = 0.0328), and IL-2 (rs = 0.3959; p = 0.0012). These results suggest that the Th17 response mediated by the IL-6 pathway has a role in the prolonged IgG response to SARS-CoV-2 infection. Full article
(This article belongs to the Special Issue Molecular Research and Insights into COVID-19 2.0)
Show Figures

Figure 1

Figure 1
<p>Graphics showing the seven cytokine profiles in the IgG response duration. The black lines represent the median values, and the dashed lines represent the quartiles.</p>
Full article ">Figure 2
<p>Dispersion plots show (<b>a</b>) the absence of correlation between plasma levels of IL-6 and IL-17A in the Short response group; and (<b>b</b>) a statistically significant correlation between plasma levels of IL-6 and IL-17A in the Long response group Dots are samples with values in the x and y axis and correlation trends are indicated by dash lines. Additionally, among the SR subgroup, no statistically significant Spearman correlations were observed between the plasma levels of IL-17A and IFN-γ (rs = 0.2399; <span class="html-italic">p</span> = 0.1043), IL-4 (rs = 0.0273; <span class="html-italic">p</span> = 0.8554), and IL-2 (rs = 0.2204; <span class="html-italic">p</span> = 0.1365). Furthermore, among the LR subgroup, weaker but statistically significant Spearman correlations were observed between the plasma levels of IL-17A and IFN-γ (rs = 0.3873; <span class="html-italic">p</span> = 0.0016), IL-4 (rs = 0.2671; <span class="html-italic">p</span> = 0.0328), and IL-2 (rs = 0.3959; <span class="html-italic">p</span> = 0.0012).</p>
Full article ">Figure 3
<p>Total plasmatic levels of total IgG, IgA, and IgM among SR and LR subgroups of patients. Black lines represent median values, and dashed lines represent quartiles.</p>
Full article ">Figure 4
<p>Description of sample design and data sources. LR = Long response subgroup; SR = Short response subgroup. The sources of data were previous studies by De Oliveira et al. [<a href="#B12-ijms-25-08740" class="html-bibr">12</a>], Soares et al. [<a href="#B11-ijms-25-08740" class="html-bibr">11</a>], Silva et al. [<a href="#B22-ijms-25-08740" class="html-bibr">22</a>], and Queiroz et al. [<a href="#B21-ijms-25-08740" class="html-bibr">21</a>]. Additionally, in the same samples, we obtained data on the plasmatic levels of the total IgG, IgA, and IgM measured by an ELISA, along with genotypic data of four SNPs (rs2228145/<span class="html-italic">IL6R</span>; rs3087456/<span class="html-italic">CIITA</span>; rs3077/<span class="html-italic">HLA-DPA1</span>; and rs9277534/<span class="html-italic">HLA-DPB1</span>) in the Belém representative population that was not reported by Silva et al. [<a href="#B22-ijms-25-08740" class="html-bibr">22</a>].</p>
Full article ">
12 pages, 567 KiB  
Conference Report
Conference Report: LPMHealthcare Emerging Viruses 2023 (EVOX23): Pandemics—Learning from the Past and Present to Prepare for the Future
by Fern Jenkins, Tobias Mapulanga, Gauri Thapa, Kelly A. S. da Costa and Nigel J. Temperton
Pathogens 2024, 13(8), 679; https://doi.org/10.3390/pathogens13080679 (registering DOI) - 10 Aug 2024
Viewed by 210
Abstract
The emergence of SARS-CoV-2 has meant that pandemic preparedness has become a major focus of the global scientific community. Gathered in the historic St Edmund Hall college in Oxford, the one-day LPMHealthcare conference on emerging viruses (6 September 2023) sought to review and [...] Read more.
The emergence of SARS-CoV-2 has meant that pandemic preparedness has become a major focus of the global scientific community. Gathered in the historic St Edmund Hall college in Oxford, the one-day LPMHealthcare conference on emerging viruses (6 September 2023) sought to review and learn from past pandemics—the current SARS-CoV-2 pandemic and the Mpox outbreak—and then look towards potential future pandemics. This includes an emphasis on monitoring the “traditional” reservoirs of viruses with zoonotic potential, as well as possible new sources of spillover events, e.g., bats, which we are coming into closer contact with due to climate change and the impacts of human activities on habitats. Continued vigilance and investment into creative scientific solutions is required for issues including the long-term physical and psychological effects of COVID-19, i.e., long COVID. The evaluation of current systems, including environmental monitoring, communication (with the public, regulatory authorities, and governments), and training; assessment of the effectiveness of the technologies/assays we have in place currently; and lobbying of the government and the public to work with scientists are all required in order to build trust moving forward. Overall, the SARS-CoV-2 pandemic has shown how many sectors can work together to achieve a global impact in times of crisis. Full article
(This article belongs to the Section Emerging Pathogens)
Show Figures

Figure 1

Figure 1
<p>Schematic showing latest emerging virus R&amp;D which builds successively on previous epidemics and pandemics. Figure generated using BioRender.</p>
Full article ">
11 pages, 1507 KiB  
Communication
Insights on the Mechanical Properties of SARS-CoV-2 Particles and the Effects of the Photosensitizer Hypericin
by Matteo Mariangeli, Ana Moreno, Pietro Delcanale, Stefania Abbruzzetti, Alberto Diaspro, Cristiano Viappiani and Paolo Bianchini
Int. J. Mol. Sci. 2024, 25(16), 8724; https://doi.org/10.3390/ijms25168724 (registering DOI) - 10 Aug 2024
Viewed by 344
Abstract
SARS-CoV-2 is a highly pathogenic virus responsible for the COVID-19 disease. It belongs to the Coronaviridae family, characterized by a phospholipid envelope, which is crucial for viral entry and replication in host cells. Hypericin, a lipophilic, naturally occurring photosensitizer, was reported to effectively [...] Read more.
SARS-CoV-2 is a highly pathogenic virus responsible for the COVID-19 disease. It belongs to the Coronaviridae family, characterized by a phospholipid envelope, which is crucial for viral entry and replication in host cells. Hypericin, a lipophilic, naturally occurring photosensitizer, was reported to effectively inactivate enveloped viruses, including SARS-CoV-2, upon light irradiation. In addition to its photodynamic activity, Hyp was found to exert an antiviral action also in the dark. This study explores the mechanical properties of heat-inactivated SARS-CoV-2 viral particles using Atomic Force Microscopy (AFM). Results reveal a flexible structure under external stress, potentially contributing to the virus pathogenicity. Although the fixation protocol causes damage to some particles, correlation with fluorescence demonstrates colocalization of partially degraded virions with their genome. The impact of hypericin on the mechanical properties of the virus was assessed and found particularly relevant in dark conditions. These preliminary results suggest that hypericin can affect the mechanical properties of the viral envelope, an effect that warrants further investigation in the context of antiviral therapies. Full article
(This article belongs to the Section Molecular Microbiology)
Show Figures

Figure 1

Figure 1
<p>Image showing a complete viral particle on a mica surface. Orange arrows point to potential remnants of spike proteins. Scale bar = 50 nm.</p>
Full article ">Figure 2
<p>Representative fatigue experiment on SARS-CoV-2. The height profiles along the green line for the first (<b>a</b>) and the last (<b>b</b>) images are plotted in panel (<b>c</b>), black and red respectively. Scale bar = 50 nm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Confocal and (<b>b</b>) AFM images of the same viral particles and (<b>c</b>) detailed image of the particle in the boxed area of panel b, that appears partially damaged. A Gaussian blur was applied to the confocal image, σ = 2. The AFM images were acquired in tapping mode. Scale bar = 200 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>) QI<sup>TM</sup> image of a few viral particles. A dashed square indicates the particle used for the breakthrough measurement reported in (<b>b</b>). Such graph displays the acquired force–distance curve showing visible penetration of the external viral envelope; (<b>c</b>) breakthrough force distributions under three different conditions with [Hyp] = 1 μM, asterisks indicate <span class="html-italic">p</span> value &lt; 0.001.</p>
Full article ">
16 pages, 6013 KiB  
Article
COVID-19 Vaccine Effectiveness Studies against Symptomatic and Severe Outcomes during the Omicron Period in Four Countries in the Eastern Mediterranean Region
by Manuela Runge, Zahra Karimian, Mehrnaz Kheirandish, Giulio Borghi, Natalie Wodniak, Kamal Fahmy, Carsten Mantel, Thomas Cherian, Zeinab Nabil Ahmed Said, Farid Najafi, Fatima Thneibat, Zia Ul-Haq, Sheraz Fazid, Iman Ibrahim Salama, Fatemeh Khosravi Shadmani, Ahmad Alrawashdeh, Shadrokh Sirous, Saverio Bellizzi, Amira Ahmed, Michael Lukwiya, Arash Rashidian and on behalf of the Consortium of Authorsadd Show full author list remove Hide full author list
Vaccines 2024, 12(8), 906; https://doi.org/10.3390/vaccines12080906 (registering DOI) - 10 Aug 2024
Viewed by 435
Abstract
Vaccine effectiveness (VE) studies provide real-world evidence to monitor vaccine performance and inform policy. The WHO Regional Office for the Eastern Mediterranean supported a regional study to assess the VE of COVID-19 vaccines against different clinical outcomes in four countries between June 2021 [...] Read more.
Vaccine effectiveness (VE) studies provide real-world evidence to monitor vaccine performance and inform policy. The WHO Regional Office for the Eastern Mediterranean supported a regional study to assess the VE of COVID-19 vaccines against different clinical outcomes in four countries between June 2021 and August 2023. Health worker cohort studies were conducted in 2707 health workers in Egypt and Pakistan, of whom 171 experienced symptomatic laboratory-confirmed SARS-CoV-2 infection. Test-negative design case–control studies were conducted in Iran and Jordan in 4017 severe acute respiratory infection (SARI) patients (2347 controls and 1670 cases) during the Omicron variant dominant period. VE estimates were calculated for each study and pooled by study design for several vaccine types (BBIBP-CorV, AZD1222, BNT162b2, and mRNA-1273, among others). Among health workers, VE against symptomatic infection of a complete primary series could only be computed compared to partial vaccination, suggesting a benefit of providing an additional dose of mRNA vaccines (VE: 88.9%, 95%CI: 15.3–98.6%), while results were inconclusive for other vaccine products. Among SARI patients, VE against hospitalization of a complete primary series with any vaccine compared to non-vaccinated was 20.9% (95%CI: 4.5–34.5%). Effectiveness estimates for individual vaccines, booster doses, and secondary outcomes (intensive care unit admission and death) were inconclusive. Future VE studies will need to address challenges in both design and analysis when conducted late during a pandemic and will be able to utilize the strengthened capacities in countries. Full article
(This article belongs to the Special Issue Immune Effectiveness of COVID-19 Vaccines)
Show Figures

Figure 1

Figure 1
<p>Sample size flowchart for the (<b>A</b>) cohort studies and (<b>B</b>) TND studies. Vaccination status in cohort studies shown at start of follow-up.</p>
Full article ">Figure 2
<p>Study population by vaccination status and time since vaccination over time. (<b>A</b>) Number of health workers in the two cohort studies by start of follow-up. The two vertical lines indicate start and end of study period. The grey shaded area indicates the pre-Omicron-dominant period. (<b>B</b>) Days since last vaccination and start of follow-up in the pooled cohort dataset (bottom) and corresponding density plot (top). The star symbol indicates outcome events during follow-up. (<b>C</b>) SARI patients in the two TND studies by admission date and vaccination status. (<b>D</b>) Days since vaccination and illness onset date in pooled TND studies (bottom) and corresponding density plot (top).</p>
Full article ">Figure 3
<p>VE of a primary series and a booster compared to partial vaccination against lab-confirmed symptomatic COVID-19 infection among health workers compared to partial vaccination (reference). Blank VE estimates indicate insufficient data to be computed.</p>
Full article ">Figure 4
<p>VE of a primary series and a booster compared to no vaccination against hospitalization among SARI patients. Blank VE fields indicate insufficient data to be computed.</p>
Full article ">Figure 5
<p>VE of a primary series and a booster compared to no vaccination against ICU admission and/or death among SARI patients in the TND studies. Blank VE estimates indicate insufficient data to be computed.</p>
Full article ">
18 pages, 1516 KiB  
Article
Post-Processing Maritime Wind Forecasts from the European Centre for Medium-Range Weather Forecasts around the Korean Peninsula Using Support Vector Regression and Principal Component Analysis
by Seung-Hyun Moon, Do-Youn Kim and Yong-Hyuk Kim
J. Mar. Sci. Eng. 2024, 12(8), 1360; https://doi.org/10.3390/jmse12081360 - 9 Aug 2024
Viewed by 209
Abstract
Accurate wind data are crucial for successful search and rescue (SAR) operations on the sea surface in maritime accidents, as survivors or debris tend to drift with the wind. As maritime accidents frequently occur outside the range of wind stations, SAR operations heavily [...] Read more.
Accurate wind data are crucial for successful search and rescue (SAR) operations on the sea surface in maritime accidents, as survivors or debris tend to drift with the wind. As maritime accidents frequently occur outside the range of wind stations, SAR operations heavily rely on wind forecasts generated by numerical models. However, numerical models encounter delays in generating results due to spin-up issues, and their predictions can sometimes exhibit inherent biases caused by geographical factors. To overcome these limitations, we reviewed the observations for the first 24 h of the 72-hour forecast from the ECMWF and then post-processed the forecast for the remaining 48 h. By effectively reducing the dimensionality of input variables comprising observation and forecast data using principal component analysis, we improved wind predictions with support vector regression. Our model achieved an average RMSE improvement of 16.01% compared to the original forecast from the ECMWF. Furthermore, it achieved an average RMSE improvement of 5.42% for locations without observation data by employing a model trained on data from the nearest wind station and then applying an adaptive weighting scheme to the output of that model. Full article
(This article belongs to the Special Issue Machine Learning Methodologies and Ocean Science)
Show Figures

Figure 1

Figure 1
<p>Locations of the 7 offshore wind stations around the Korean Peninsula.</p>
Full article ">Figure 2
<p>Illustration of PCA for two variables: <span class="html-italic">ECMWF forecast u-comp</span>. 1 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math> and <span class="html-italic">Observation u-comp</span>. 1 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math>.</p>
Full article ">Figure 3
<p>Support vector regression with epsilon tube.</p>
Full article ">Figure 4
<p>Flowchart of the wind prediction correction using PCA and SVR.</p>
Full article ">Figure 5
<p>Flowchart of the wind prediction correction using the adaptive weighting scheme.</p>
Full article ">Figure 6
<p>Monthly RMSE comparison of ECMWF and SVR preceded by PCA.</p>
Full article ">Figure 6 Cont.
<p>Monthly RMSE comparison of ECMWF and SVR preceded by PCA.</p>
Full article ">Figure 7
<p>RMSE comparison of wind prediction. ECMWF (previous day) and its post-processed version with SVR cover the 24 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math>–71 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math> prediction period, while ECMWF (current day) covers the same time period with the 0 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math>–48 <math display="inline"><semantics> <mi mathvariant="normal">h</mi> </semantics></math> forecast.</p>
Full article ">Figure A1
<p>Comparison of linear and RBF kernels in SVR.</p>
Full article ">Figure A2
<p>Comparison of MAE for wind direction prediction.</p>
Full article ">Figure A3
<p>Performance comparison based on similarity measures in adaptive weighting scheme.</p>
Full article ">
23 pages, 5374 KiB  
Article
Leveraging Visual Language Model and Generative Diffusion Model for Zero-Shot SAR Target Recognition
by Junyu Wang, Hao Sun, Tao Tang, Yuli Sun, Qishan He, Lin Lei and Kefeng Ji
Remote Sens. 2024, 16(16), 2927; https://doi.org/10.3390/rs16162927 (registering DOI) - 9 Aug 2024
Viewed by 236
Abstract
Simulated data play an important role in SAR target recognition, particularly under zero-shot learning (ZSL) conditions caused by the lack of training samples. The traditional SAR simulation method is based on manually constructing target 3D models for electromagnetic simulation, which is costly and [...] Read more.
Simulated data play an important role in SAR target recognition, particularly under zero-shot learning (ZSL) conditions caused by the lack of training samples. The traditional SAR simulation method is based on manually constructing target 3D models for electromagnetic simulation, which is costly and limited by the target’s prior knowledge base. Also, the unavoidable discrepancy between simulated SAR and measured SAR makes the traditional simulation method more limited for target recognition. This paper proposes an innovative SAR simulation method based on a visual language model and generative diffusion model by extracting target semantic information from optical remote sensing images and transforming it into a 3D model for SAR simulation to address the challenge of SAR target recognition under ZSL conditions. Additionally, to reduce the domain shift between the simulated domain and the measured domain, we propose a domain adaptation method based on dynamic weight domain loss and classification loss. The effectiveness of semantic information-based 3D models has been validated on the MSTAR dataset and the feasibility of the proposed framework has been validated on the self-built civilian vehicle dataset. The experimental results demonstrate that the first proposed SAR simulation method based on a visual language model and generative diffusion model can effectively improve target recognition performance under ZSL conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Framework of our method.</p>
Full article ">Figure 2
<p>Extracting target semantic information from optical remote sensing image.</p>
Full article ">Figure 3
<p>Target semantic information diffusion to 3D model (example image of T-72 tank from Wikimedia Commons [<a href="#B47-remotesensing-16-02927" class="html-bibr">47</a>]).</p>
Full article ">Figure 4
<p>The influence of target key features on SAR simulation.</p>
Full article ">Figure 5
<p>Simulation SAR with domain adaption for target recognition.</p>
Full article ">Figure 6
<p>The dataset used in this paper: (<b>a</b>) MSTAR, (<b>b</b>) civilian vehicle dataset.</p>
Full article ">Figure 7
<p>Dataset for fine-tuning: (<b>a</b>) military vehicle fine-tuning set, (<b>b</b>) civilian vehicle semantic extraction set, (<b>c</b>) civilian vehicle fine-tuning set.</p>
Full article ">Figure 8
<p>Generating 3D models based on target semantic information.</p>
Full article ">Figure 9
<p>Comparison between simulated SAR and measured SAR, the target and shadow are circled with red lines respectively.</p>
Full article ">Figure 10
<p>Cosine similarity between simulated and measured images. (<b>a</b>) SAMPLE. (<b>b</b>) Our simulated SAR.</p>
Full article ">Figure 11
<p>Confusion matrix for five types of military targets. (<b>a</b>) SAMPLE simulated SAR direct training. (<b>b</b>) Our simulated SAR direct training. (<b>c</b>) SAMPLE simulated SAR with DANN-<math display="inline"><semantics> <msub> <mi>W</mi> <mi>n</mi> </msub> </semantics></math>. (<b>d</b>) Our simulated SAR with DANN-<math display="inline"><semantics> <msub> <mi>W</mi> <mi>n</mi> </msub> </semantics></math>.</p>
Full article ">Figure 12
<p>t-SNE of five types of military targets. (<b>a</b>) SAMPLE simulated SAR direct training. (<b>b</b>) SAMPLE simulated SAR with DANN-<math display="inline"><semantics> <msub> <mi>W</mi> <mi>n</mi> </msub> </semantics></math>.</p>
Full article ">Figure 13
<p>Generating 3D models of civilian vehicles using target semantic information.</p>
Full article ">Figure 14
<p>A comparison of the measured and simulated SAR of civilian vehicles (the top column is the measured image, and the bottom column is the simulated image).</p>
Full article ">Figure 15
<p>Confusion matrix of SAR civilian vehicle target recognition. (<b>a</b>) DANN-<math display="inline"><semantics> <msub> <mi>W</mi> <mi>n</mi> </msub> </semantics></math>. (<b>b</b>) Confusion matrix of ConvNeXt-T. (<b>c</b>) Confusion matrix of Vgg19. (<b>d</b>) Confusion matrix of AlexNet. (<b>e</b>) Confusion matrix of ResNet50.</p>
Full article ">
Back to TopTop