[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 (69)

Search Parameters:
Keywords = VPM

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
7 pages, 696 KiB  
Proceeding Paper
Using SABC Algorithm for Scheduling Unrelated Parallel Batch Processing Machines Considering Deterioration Effects and Variable Maintenance
by Ziyang Ji, Jabir Mumtaz and Ke Ke
Eng. Proc. 2024, 75(1), 20; https://doi.org/10.3390/engproc2024075020 - 24 Sep 2024
Viewed by 156
Abstract
This paper investigates the problem of processing jobs on unrelated parallel batch machines, taking into account job arrival times, machine deterioration effects, and variable preventive maintenance (VPM). To address this complex scheduling problem, this paper proposes a Self-Adaptive Artificial Bee Colony (SABC) algorithm, [...] Read more.
This paper investigates the problem of processing jobs on unrelated parallel batch machines, taking into account job arrival times, machine deterioration effects, and variable preventive maintenance (VPM). To address this complex scheduling problem, this paper proposes a Self-Adaptive Artificial Bee Colony (SABC) algorithm, incorporating an adaptive variable neighborhood search mechanism into the algorithm. To verify the effectiveness of the proposed algorithm, we designed comparative experiments, comparing the SABC algorithm with the NSGA-III algorithm on problem instances of different scales. The results indicate that the SABC algorithm outperforms the NSGA-III algorithm in terms of solution quality and diversity, and this advantage becomes more pronounced as the problem scale increases. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the proposed SABC.</p>
Full article ">Figure 2
<p>Example of neighborhood structure.</p>
Full article ">
27 pages, 6924 KiB  
Article
GPP of a Chinese Savanna Ecosystem during Different Phenological Phases Simulated from Harmonized Landsat and Sentinel-2 Data
by Xiang Zhang, Shuai Xie, Yiping Zhang, Qinghai Song, Gianluca Filippa and Dehua Qi
Remote Sens. 2024, 16(18), 3475; https://doi.org/10.3390/rs16183475 - 19 Sep 2024
Viewed by 825
Abstract
Savannas are widespread biomes with highly valued ecosystem services. To successfully manage savannas in the future, it is critical to better understand the long-term dynamics of their productivity and phenology. However, accurate large-scale gross primary productivity (GPP) estimation remains challenging because of the [...] Read more.
Savannas are widespread biomes with highly valued ecosystem services. To successfully manage savannas in the future, it is critical to better understand the long-term dynamics of their productivity and phenology. However, accurate large-scale gross primary productivity (GPP) estimation remains challenging because of the high spatial and seasonal variations in savanna GPP. China’s savanna ecosystems constitute only a small part of the world’s savanna ecosystems and are ecologically fragile. However, studies on GPP and phenological changes, while closely related to climate change, remain scarce. Therefore, we simulated savanna ecosystem GPP via a satellite-based vegetation photosynthesis model (VPM) with fine-resolution harmonized Landsat and Sentinel-2 (HLS) imagery and derived savanna phenophases from phenocam images. From 2015 to 2018, we compared the GPP from HLS VPM (GPPHLS-VPM) simulations and that from Moderate-Resolution Imaging Spectroradiometer (MODIS) VPM simulations (GPPMODIS-VPM) with GPP estimates from an eddy covariance (EC) flux tower (GPPEC) in Yuanjiang, China. Moreover, the consistency of the savanna ecosystem GPP was validated for a conventional MODIS product (MOD17A2). This study clearly revealed the potential of the HLS VPM for estimating savanna GPP. Compared with the MODIS VPM, the HLS VPM yielded more accurate GPP estimates with lower root-mean-square errors (RMSEs) and slopes closer to 1:1. Specifically, the annual RMSE values for the HLS VPM were 1.54 (2015), 2.65 (2016), 2.64 (2017), and 1.80 (2018), whereas those for the MODIS VPM were 3.04, 3.10, 2.62, and 2.49, respectively. The HLS VPM slopes were 1.12, 1.80, 1.65, and 1.27, indicating better agreement with the EC data than the MODIS VPM slopes of 2.04, 2.51, 2.14, and 1.54, respectively. Moreover, HLS VPM suitably indicated GPP dynamics during all phenophases, especially during the autumn green-down period. As the first study that simulates GPP involving HLS VPM and compares satellite-based and EC flux observations of the GPP in Chinese savanna ecosystems, our study enables better exploration of the Chinese savanna ecosystem GPP during different phenophases and more effective savanna management and conservation worldwide. Full article
(This article belongs to the Special Issue Remote Sensing of Savannas and Woodlands II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area and the Yuanjiang (YJ) savanna flux tower site.</p>
Full article ">Figure 2
<p>Workflow of this study. The light blue rectangular boxes denote the key processes, and the dark green rectangular boxes denote the important data and result outputs.</p>
Full article ">Figure 3
<p>Variations in temperature (<b>a</b>) and precipitation (<b>b</b>) at the Yuanjiang station from 2015 to 2018.</p>
Full article ">Figure 4
<p>Time series of the daily GCC derived from phenocam images and HLS-based VIs (NDVI, LSWI, and EVI) at the Yuanjiang savanna site from September 2015 to December 2018. Because the Sentinel-2 satellite was launched in June 2015, the HLS time series data began in September 2015 after quality control.</p>
Full article ">Figure 5
<p>Daily GCC derived from phenocam images and MODIS-based VIs (NDVI, LSWI, and EVI) at the Yuanjiang savanna site from 2015 to 2018.</p>
Full article ">Figure 6
<p>Linear correlation coefficients between the HLS- (<b>a</b>–<b>c</b>) and MODIS-based VIs (<b>d</b>–<b>f</b>) and the flux observation-based GPP<sub>EC</sub>.</p>
Full article ">Figure 7
<p>Comparison between the interannual variations in the EC tower-derived gross primary production (GPP<sub>EC</sub>) and the simulated GPP (GPPvpm) from the two remote sensing data sources: (<b>a</b>) MODIS and (<b>b</b>) HLS.</p>
Full article ">Figure 8
<p>Linear comparisons of the tower-based gross primary production (EC) with the VPM GPP estimates for the HLS (HLS VPM) (<b>a</b>–<b>d</b>) and MODIS (MODIS VPM) data (<b>e</b>–<b>h</b>) from 2015 to 2018. R<sup>2</sup>: coefficient of determination; the fitting equation for y and x and 0 intercepts are provided.</p>
Full article ">Figure 9
<p>Comparison between the interannual variations in the tower-derived gross primary production (GPP<sub>EC</sub>) and the MODIS (MOD17A2)-based GPP.</p>
Full article ">Figure 10
<p>Example of a Yuanjiang savanna photograph processed with the xROI package (<b>a</b>) and 2015–2018 time series of the daily GCC in the Yuanjiang savanna (<b>b</b>). The three red masks denoting the ROI (<b>a</b>). The photo dates are 8 June 2015, and the file name is yj_2015_06_08_120412. The JPG image follows the phenocam convention. The GCC index is a dimensionless, calculated value extracted from digital photographs over the 2015–2018 period (<b>b</b>).</p>
Full article ">Figure 11
<p>Linear regression equations with an intercept of 0 for the HLS VPM simulations compared with the EC data during the different phenophases from 2015 to 2018 (<b>a</b>–<b>d</b>) for the green-up phenophase of each year and (<b>e</b>–<b>h</b>) for the green-down phenophase of each year.</p>
Full article ">Figure 12
<p>Linear regression equations with an intercept of 0 for the MODIS VPM GPP simulations compared with the EC data during the different phenophases from 2015 to 2018 (<b>a</b>–<b>d</b>) for the green-up phenophase of each year and (<b>e</b>–<b>h</b>) for the green-down phenophase of each year.</p>
Full article ">Figure 13
<p>Linear regression equations with an intercept of 0 for the MOD17A2 GPP simulations compared with the EC data during the different phenophases from 2015 to 2018 (<b>a</b>–<b>d</b>) for the green-up phenophase of each year and (<b>e</b>–<b>h</b>) for the green-down phenophase of each year.</p>
Full article ">
21 pages, 13840 KiB  
Article
Estimating Forest Gross Primary Production Using Machine Learning, Light Use Efficiency Model, and Global Eddy Covariance Data
by Zhenkun Tian, Yingying Fu, Tao Zhou, Chuixiang Yi, Eric Kutter, Qin Zhang and Nir Y. Krakauer
Forests 2024, 15(9), 1615; https://doi.org/10.3390/f15091615 - 13 Sep 2024
Viewed by 604
Abstract
Forests play a vital role in atmospheric CO2 sequestration among terrestrial ecosystems, mitigating the greenhouse effect induced by human activity in a changing climate. The LUE (light use efficiency) model is a popular algorithm for calculating terrestrial GPP (gross primary production) based [...] Read more.
Forests play a vital role in atmospheric CO2 sequestration among terrestrial ecosystems, mitigating the greenhouse effect induced by human activity in a changing climate. The LUE (light use efficiency) model is a popular algorithm for calculating terrestrial GPP (gross primary production) based on physiological mechanisms and is easy to implement. Different versions have been applied for many years to simulate the GPP of different ecosystem types at regional or global scales. For estimating forest GPP using different approaches, we implemented five LUE models (EC-LUE, VPM, GOL-PEM, CASA, and C-Fix) in forests of type DBF, EBF, ENF, and MF, using the FLUXNET2015 dataset, remote sensing observations, and Köppen–Geiger climate zones. We then fused these models to additionally improve the ability of the GPP estimation using an RF (random forest) and an SVM (support vector machine). Our results indicated that under a unified parameterization scheme, EC-LUE and VPM yielded the best performance in simulating GPP variations, followed by GLO-PEM, CASA, and C-fix, while MODIS also demonstrated reliable GPP estimation ability. The results of the model fusion across different forest types and flux net sites indicated that the RF could capture more GPP variation magnitudes with higher R2 and lower RMSE than the SVM. Both RF and SVM were validated using cross-validation for all forest types and flux net sites, showing that the accuracy of the GPP simulation could be improved by the RF and SVM by 28% and 27%. Full article
(This article belongs to the Section Forest Ecology and Management)
Show Figures

Figure 1

Figure 1
<p>Köppen–Geiger climate zones and 45 FLUXNET2015 forest sites (red triangles) distribution. Köppen–Geiger climate symbols are listed in <a href="#app1-forests-15-01615" class="html-app">Table S2 in the Supporting Information File</a>.</p>
Full article ">Figure 2
<p>Workflow of GPP estimation through the integration of LUE models based on ground measurements, remote sensing observations, and Köppen–Geiger climate zones.</p>
Full article ">Figure 3
<p>The Taylor diagrams for site-derived GPP and LUE models/machine learning estimates at the 45 FLUXNET2015 sites. The dotted circular lines which connect the X and Y axes denote <span class="html-italic">SD</span>. The dotted radial lines represent R. The brown curves are <span class="html-italic">RMSD</span> compared to the referenced site’s GPP.</p>
Full article ">Figure 4
<p>The <span class="html-italic">R</span><sup>2</sup> (<b>a</b>), <span class="html-italic">RMSE</span> (<b>b</b>), and <span class="html-italic">RPE</span> (<b>c</b>) of 5 single models, MODIS, SVM, and RF across the DBF, EBF, ENF, and MF.</p>
Full article ">Figure 5
<p>The scatter plots of <span class="html-italic">R</span><sup>2</sup>, <span class="html-italic">RMSE</span>, and <span class="html-italic">RPE</span> across the DBF between site-derived GPP and the estimates from the LUE models, MODIS, SVM, and RF.</p>
Full article ">Figure 6
<p>The probability distribution of errors from the LUE models, MODIS, SVM, and RF.</p>
Full article ">Figure 7
<p>The <span class="html-italic">AIC</span> (<b>a</b>) and <span class="html-italic">BIC</span> (<b>b</b>) of the LUE models, MODIS, SVM, and RF across the DBF, EBF, ENF, and MF.</p>
Full article ">Figure 8
<p>Daily FLUXNET2015’s GPP (black dots), the best LUE model-estimated GPP (EC-LUE, line in blue), and the best fusion method-estimated GPP (RF, line in orange) at 4 sites: DE-Hai of DBF (<b>a</b>), AU-Tum of EBF (<b>b</b>), US-Blo of ENF (<b>c</b>), and BE-Bra of MF (<b>d</b>).</p>
Full article ">Figure 9
<p>Boxplot of performance of <span class="html-italic">R</span><sup>2</sup> (<b>a</b>), <span class="html-italic">RMSE</span> (<b>b</b>), and <span class="html-italic">AIC</span> (<b>c</b>) of LUE and machine learning methods across 45 FLUXNET2015 sites.</p>
Full article ">
14 pages, 5003 KiB  
Article
Production, Passaging Stability, and Histological Analysis of Madin–Darby Canine Kidney Cells Cultured in a Low-Serum Medium
by Ming Cai, Yang Le, Zheng Gong, Tianbao Dong, Bo Liu, Minne Su, Xuedan Li, Feixia Peng, Qingda Li, Xuanxuan Nian, Hao Yu, Zheng Wu, Zhegang Zhang and Jiayou Zhang
Vaccines 2024, 12(9), 991; https://doi.org/10.3390/vaccines12090991 - 30 Aug 2024
Viewed by 562
Abstract
Madin–Darby canine kidney (MDCK) cells are commonly used to produce cell-based influenza vaccines. However, the role of the low-serum medium on the proliferation of MDCK cells and the propagation of the influenza virus has not been well studied. In the present study, we [...] Read more.
Madin–Darby canine kidney (MDCK) cells are commonly used to produce cell-based influenza vaccines. However, the role of the low-serum medium on the proliferation of MDCK cells and the propagation of the influenza virus has not been well studied. In the present study, we used 5 of 15 culture methods with different concentrations of a mixed medium and neonatal bovine serum (NBS) to determine the best culture medium. We found that a VP:M199 ratio of 1:2 (3% NBS) was suitable for culturing MDCK cells. Furthermore, the stable growth of MDCK cells and the production of the influenza virus were evaluated over long-term passaging. We found no significant difference in terms of cell growth and virus production between high and low passages of MDCK cells under low-serum culture conditions, regardless of influenza virus infection. Lastly, we performed a comparison of the transcriptomics and proteomics of MDCK cells cultured in VP:M199 = 1:2 (3% NBS) with those cultured in VP:M199 = 1:2 (5% NBS) before and after influenza virus infection. The transcriptome analysis showed that differentially expressed genes were predominantly enriched in the metabolic pathway and MAPK signaling pathway, indicating an activated state. This suggests that decreasing the concentration of serum in the medium from 5% to 3% may increase the metabolic activity of cells. Proteomics analysis showed that only a small number of differentially expressed proteins could not be enriched for analysis, indicating minimal difference in the protein levels of MDCK cells when the serum concentration in the medium was decreased from 5% to 3%. Altogether, our findings suggest that the screening and application of a low-serum medium provide a background for the development and optimization of cell-based influenza vaccines. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Growth curves of MDCK cells in 15 cultures; VP:M199 = 1:2 (3%NBS, 2% NBS), VP:M199 = 1:3 (4% NBS, 3% NBS), and VP:M199 = 1:4 (3% NBS) cultures were selected for the validation of cell factory scale-up cultures (blue: control, red: experimental).</p>
Full article ">Figure 2
<p>MDCK cell factories (CF1, CF2, CF10) enlarged culture validation results. ns Not Significant, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>MDCK cell culture continuous passaging stability ((<b>A</b>) cell growth status, (<b>B</b>) cell metabolism). ns Not Significant.</p>
Full article ">Figure 4
<p>Growth curves of MDCK cells of different generations (<b>A</b>–<b>D</b>) Solid line: cell density, dashed line: cell viability and cell colony formation (<b>E</b>,<b>F</b>); numbered as serum content—cell generation—inoculum density, example: 3 (3% NBS)—65 (Cell generation)—1 (Inoculated at a density of 5 × 10<sup>4</sup> cells/mL). ns Not Significant.</p>
Full article ">Figure 5
<p>Effect of low-serum medium on the viral susceptibility of different generations of MDCK cells ((<b>A</b>) H1N1 virus susceptibility, (<b>B</b>) H3N2 virus susceptibility, (<b>C</b>) BV virus susceptibility, (<b>D</b>): BY virus susceptibility). (3% NBS)—65: (Serum content)-cell generation. ns Not Significant.</p>
Full article ">Figure 6
<p>Results of transcriptomics and proteomics analyses. (<b>A</b>) Volcano plot showing DEGs of 3-uninfected group vs. 5-uninfected group. Red, blue, and black dots stand for genes with upregulation, downregulation, and non-differentiation, respectively. (<b>B</b>) Volcano plot showing DEGs of 3-infected group vs. 5-infected group. (<b>C</b>) Dot plot of 3-uninfected group vs. 5-uninfected group KEGG enrichment. (<b>D</b>) Dot plot of 3-infected group vs. 5-infected group KEGG enrichment. (<b>E</b>) Heatmap showing differentially expressed proteins between the 3-uninfected group and the 5-uninfected group. (<b>F</b>) Heatmap showing differentially expressed proteins between the 3-infected group and the 5-infected group. (<b>G</b>) Venn diagram showing the results of transcriptomic and proteomic correlation analysis between the 3-uninfected group and the 5-uninfected group. (<b>H</b>) Venn diagram showing the results of transcriptomic and proteomic correlation analysis between the 3-infected group and the 5-infected group.</p>
Full article ">
28 pages, 2798 KiB  
Article
An rVPM-Based Aerodynamic Hybrid Optimization Method for Coaxial Rotor with Differentiated Upper and Lower Blades in Both Hover and High-Speed Cruising States
by Zhiwei Ding, Dengyan Duan, Chaoqun Zhang and Jianbo Li
Aerospace 2024, 11(6), 463; https://doi.org/10.3390/aerospace11060463 - 9 Jun 2024
Viewed by 712
Abstract
To enhance the performance of rigid coaxial rotors across both hovering and high-speed cruising conditions, this study develops a novel aerodynamic optimization method that differentiates between the upper and lower rotors. Utilizing the lifting line and reformulated viscous vortex particle method (rVPM), this [...] Read more.
To enhance the performance of rigid coaxial rotors across both hovering and high-speed cruising conditions, this study develops a novel aerodynamic optimization method that differentiates between the upper and lower rotors. Utilizing the lifting line and reformulated viscous vortex particle method (rVPM), this approach models the complex wake fields of coaxial rotors and accurately assesses the aerodynamic loads on the blades. The optimization of geometric properties such as planform configuration and nonlinear twist is conducted through an innovative solver that integrates simulated annealing with the Nelder–Mead algorithm, ensuring both rapid and comprehensive optimization results. Comparative analyses demonstrate that these tailored geometric adjustments significantly enhance efficiency in both operational states, surpassing traditional methods. This research provides a strategic framework for addressing the varied aerodynamic challenges presented by different flight states in coaxial rotor design. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>Schematic of vortex particle generation in coaxial rotors.</p>
Full article ">Figure 2
<p>Comparison of a scaled X2TD-like rotor performance in hover and forward flight: (<b>a</b>) the figure of merit of hover at different blade loadings, and (<b>b</b>) the equivalent rotor lift-to-drag ratio at different advance ratios.</p>
Full article ">Figure 3
<p>Validation of the required power of X2TD.</p>
Full article ">Figure 4
<p>Fitting curve of rotor blade chord distribution of X2TD based on B-spline method.</p>
Full article ">Figure 5
<p>Optimization process of SANM hybrid strategy.</p>
Full article ">Figure 6
<p>Objective function value comparison during optimization process.</p>
Full article ">Figure 7
<p>Vorticity wake of rigid coaxial rotor in (<b>a</b>) hover and (<b>b</b>) <math display="inline"><semantics> <mi>μ</mi> </semantics></math> = 0.6 cruise states.</p>
Full article ">Figure 8
<p>Schematic diagram of Pareto front solution set.</p>
Full article ">Figure 9
<p>Planform of optimized blade and X2TD blade.</p>
Full article ">Figure 10
<p>Distributionof chord and twist of optimized blade and XT2D blade.</p>
Full article ">Figure 11
<p>Hover performance changes with thrust coefficient.</p>
Full article ">Figure 12
<p>Thrust coefficient distribution spanwise.</p>
Full article ">Figure 13
<p>Angle of attack distribution spanwise.</p>
Full article ">Figure 14
<p>Cruise performance comparison of optimized and X2TD blade.</p>
Full article ">Figure 15
<p>Thrust coefficient distribution spanwise of upper rotor (<math display="inline"><semantics> <mi>ψ</mi> </semantics></math> = 90°, <math display="inline"><semantics> <mi>μ</mi> </semantics></math> = 0.35).</p>
Full article ">Figure 16
<p>Thrust coefficient distribution spanwise of upper rotor (<math display="inline"><semantics> <mi>ψ</mi> </semantics></math> = 90°, <math display="inline"><semantics> <mi>μ</mi> </semantics></math> = 0.5).</p>
Full article ">Figure 17
<p>Thrust coefficient distribution spanwise of upper rotor (<math display="inline"><semantics> <mi>ψ</mi> </semantics></math> = 90°, <math display="inline"><semantics> <mi>μ</mi> </semantics></math> = 0.6).</p>
Full article ">Figure 18
<p>Tip vortex structure of upper rotor at different advance ratios.</p>
Full article ">Figure 19
<p>Angle of attack distribution on the rotor disk (<math display="inline"><semantics> <mi>μ</mi> </semantics></math> = 0.6).</p>
Full article ">
13 pages, 8071 KiB  
Article
Biosecurity Insights from the United States Swine Health Improvement Plan: Analyzing Data to Enhance Industry Practices
by Michael Harlow, Montserrat Torremorell, Cristopher J. Rademacher, Jordan Gebhardt, Tyler Holck, Leticia C. M. Linhares, Rodger G. Main and Giovani Trevisan
Animals 2024, 14(7), 1134; https://doi.org/10.3390/ani14071134 - 8 Apr 2024
Viewed by 1897
Abstract
Biosecurity practices aim to reduce the frequency of disease outbreaks in a farm, region, or country and play a pivotal role in fortifying the country’s pork industry against emerging threats, particularly foreign animal diseases (FADs). This article addresses the current biosecurity landscape of [...] Read more.
Biosecurity practices aim to reduce the frequency of disease outbreaks in a farm, region, or country and play a pivotal role in fortifying the country’s pork industry against emerging threats, particularly foreign animal diseases (FADs). This article addresses the current biosecurity landscape of the US swine industry by summarizing the biosecurity practices reported by the producers through the United States Swine Health Improvement Plan (US SHIP) enrollment surveys, and it provides a general assessment of practices implemented. US SHIP is a voluntary, collaborative effort between industry, state, and federal entities regarding health certification programs for the swine industry. With 12,195 sites surveyed across 31 states, the study provides a comprehensive snapshot of current biosecurity practices. Key findings include variability by site types that have completed Secure Pork Supply plans, variability in outdoor access and presence of perimeter fencing, and diverse farm entry protocols for visitors. The data also reflect the industry’s response to the threat of FADs, exemplified by the implementation of the US SHIP in 2020. As the US SHIP program advances, these insights will guide industry stakeholders in refining biosecurity practices, fostering endemic re-emerging and FAD preparedness, and ensuring the sustainability of the swine industry in the face of evolving challenges. Full article
(This article belongs to the Special Issue Biosecuring Animal Populations)
Show Figures

Figure 1

Figure 1
<p>Number of sites by type of production site.</p>
Full article ">Figure 2
<p>Number of sites that completed the US SHIP survey at enrollment by state.</p>
Full article ">Figure 3
<p>Percentage of sites by site type that have completed the Secure Pork Supply (SPS) plans.</p>
Full article ">Figure 4
<p>Percent of sites by site type where animals have access to the outdoors.</p>
Full article ">Figure 5
<p>Percentage of sites, by site type, where sites have perimeter fences.</p>
Full article ">Figure 6
<p>Percentage of responses by procedure frequency where visitors sign a log-in book prior to entering a site(s).</p>
Full article ">Figure 7
<p>Percentage of sites that require a given procedure to enter a farm.</p>
Full article ">Figure 8
<p>Frequency of feed mitigants used in feed rations to reduce disease transmission risk.</p>
Full article ">Figure 9
<p>Frequency that feed supplier(s) have held imported feed ingredients to reduce disease transmission risk.</p>
Full article ">Figure 10
<p>Frequency of transport trailers reported to have been washed before returning to the point of concentration broken down by type of production site.</p>
Full article ">
28 pages, 5129 KiB  
Article
Numerical Evaluation of Aircraft Aerodynamic Static and Dynamic Stability Derivatives by a Mid-Fidelity Approach
by Daniele Granata, Alberto Savino and Alex Zanotti
Aerospace 2024, 11(3), 213; https://doi.org/10.3390/aerospace11030213 - 8 Mar 2024
Cited by 1 | Viewed by 1959
Abstract
The present study aimed to investigate the capability of mid-fidelity aerodynamic solvers in performing a preliminary evaluation of the static and dynamic stability derivatives of aircraft configurations in their design phase. In this work, the mid-fidelity aerodynamic solver DUST, which is based [...] Read more.
The present study aimed to investigate the capability of mid-fidelity aerodynamic solvers in performing a preliminary evaluation of the static and dynamic stability derivatives of aircraft configurations in their design phase. In this work, the mid-fidelity aerodynamic solver DUST, which is based on the novel vortex particle method (VPM), was used to perform simulations of the static and dynamic motion conditions of the Stability And Control CONfiguration (SACCON): an unmanned combat aerial vehicle geometry developed by NATO’s Research and Technology Organisation (RTO), which is used as a benchmark test case in the literature for the evaluation of aircraft stability derivatives. Two different methods were exploited to extract the dynamic stability derivative values from the aerodynamic coefficient time histories that were calculated with DUST. The results for the mid-fidelity approach were in good agreement with the obtained experimental data, as well as with the results obtained using more demanding high-fidelity CFD simulations. This demonstrates its suitability when implemented in DUST for predicting the static and dynamic behavior of airloads in different conditions, as well as in reliably predicting the values of stability derivatives, with the advantage of requiring limited computational effort with respect to classical high-fidelity numerical approaches and the use of wind tunnel tests. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>Graphical representation of the extraction of the dynamic derivative.</p>
Full article ">Figure 2
<p>Graphical representation of the rolling motion.</p>
Full article ">Figure 3
<p>Graphical representation of the pitching motion.</p>
Full article ">Figure 4
<p>Graphical representation of the yawing motion.</p>
Full article ">Figure 5
<p>Graphical representation of the phugoid motion.</p>
Full article ">Figure 6
<p>Graphical representation of the plunging motion.</p>
Full article ">Figure 7
<p>(<b>a</b>) SACCON geometrical scheme, the moment reference point (MRP), and the point of rotation (picture from [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]). (<b>b</b>) Definitions of the aerodynamic coefficients and reference frames.</p>
Full article ">Figure 8
<p>(<b>a</b>) Results of the mesh discretization sensitivity analysis. (<b>b</b>) Final mesh of the SACCON aircraft.</p>
Full article ">Figure 9
<p>Results of the box length sensitivity analysis.</p>
Full article ">Figure 10
<p>Results of the time discretization analysis for the aircraft rolling motion.</p>
Full article ">Figure 11
<p>Static stability analysis: the aerodynamic coefficients as a function of the angle of attack <math display="inline"><semantics> <mi>α</mi> </semantics></math> in comparison with the CFD <span class="html-italic">elsA</span> [<a href="#B12-aerospace-11-00213" class="html-bibr">12</a>] and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>SACCON static stability derivatives with respect to the angle of attack in comparison with the CFD <span class="html-italic">elsA</span> [<a href="#B12-aerospace-11-00213" class="html-bibr">12</a>] and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 13
<p>Pressure coefficient comparison that was evaluated at <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 10° on the aircraft longitudinal sections defined in [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>] in comparison with the CFD <span class="html-italic">elsA</span> [<a href="#B12-aerospace-11-00213" class="html-bibr">12</a>] and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 14
<p>Streamline visualizations and pressure coefficient contours that were evaluated at <math display="inline"><semantics> <mi>α</mi> </semantics></math> = <math display="inline"><semantics> <mrow> <msup> <mn>5.28</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>, <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 10°, and <math display="inline"><semantics> <mi>α</mi> </semantics></math> = <math display="inline"><semantics> <mrow> <msup> <mn>16.83</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. Comparison between the <span class="html-italic">DUST</span> and CFD simulations [<a href="#B12-aerospace-11-00213" class="html-bibr">12</a>,<a href="#B13-aerospace-11-00213" class="html-bibr">13</a>].</p>
Full article ">Figure 15
<p>Static stability analysis: the aerodynamic coefficients as a function of the sideslip angle <math display="inline"><semantics> <mi>β</mi> </semantics></math> in comparison with the DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>SACCON static stability derivatives with respect to the sideslip angle in comparison with the DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 17
<p>Visualization of the <span class="html-italic">DUST</span> solution for the roll oscillation cycle with an amplitude of <math display="inline"><semantics> <mi>ϕ</mi> </semantics></math> = 5° around <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0°. The particles were wake-colored by singularity intensities and the contour of pressure coefficients on the wing surface. (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>°, <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0</mn> <mspace width="4pt"/> <mi>T</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mo>+</mo> <mn>5</mn> </mrow> </semantics></math>°, <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.25</mn> <mspace width="4pt"/> <mi>T</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>°, <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.5</mn> <mspace width="4pt"/> <mi>T</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics></math>°, <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.75</mn> <mspace width="4pt"/> <mi>T</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 18
<p>Roll oscillation aerodynamic coefficient time histories in comparison with the <span class="html-italic">DUST</span> simulations and NASA’s Langley experiments, T134 and Run 15 [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>43</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.126</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.40</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 19
<p>Out-of-phase rolling dynamic stability derivatives that were computed with different methods in comparison with the <span class="html-italic">DUST</span> and NASA Langley T134 and Run 15 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 20
<p>The roll oscillation aerodynamic coefficient time histories at <math display="inline"><semantics> <mrow> <msub> <mi>α</mi> <mn>0</mn> </msub> <mo>=</mo> </mrow> <msup> <mn>10</mn> <mo>∘</mo> </msup> </semantics></math> in comparison with the <span class="html-italic">DUST</span> simulations and NASA’s Langley T134 and Run 15 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>43</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.126</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.40</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 21
<p>Pitch oscillation aerodynamic coefficient time histories in comparison with the <span class="html-italic">DUST</span> simulations and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 22
<p>The out-of-phase pitching dynamic stability derivatives that were computed with different methods in comparison with the <span class="html-italic">DUST</span> and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 23
<p>The yaw oscillation aerodynamic coefficients’ time histories obtained via the <span class="html-italic">DUST</span> simulations. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 24
<p>The out-of-phase yawing dynamic stability derivatives that were computed with different methods in comparison with the <span class="html-italic">DUST</span> and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">Figure 25
<p>The longitudinal phugoid motion aerodynamic coefficients time histories that were computed with <span class="html-italic">DUST</span>. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 26
<p>The SACCON separated pitch rate dynamic stability derivative in comparison with the <span class="html-italic">DUST</span> simulations and high-fidelity CFD <span class="html-italic">elsA</span> experiments [<a href="#B12-aerospace-11-00213" class="html-bibr">12</a>].</p>
Full article ">Figure 27
<p>The plunging motion aerodynamic coefficient time histories in comparison with the <span class="html-italic">DUST</span> and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>]. <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mo movablelimits="true" form="prefix">inf</mo> </msub> <mo>=</mo> <mn>50</mn> <mo> </mo> <mrow> <mi mathvariant="normal">m</mi> <mo> </mo> <msup> <mi mathvariant="normal">s</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>M</mi> <mi>a</mi> <mo>=</mo> <mn>0.147</mn> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>R</mi> <mi>e</mi> <mo>=</mo> <mn>1.63</mn> <mo>×</mo> <msup> <mn>10</mn> <mn>6</mn> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 28
<p>The SACCON separated <math display="inline"><semantics> <mover accent="true"> <mi>α</mi> <mo>˙</mo> </mover> </semantics></math> dynamic stability derivatives in comparison with the <span class="html-italic">DUST</span> and DNW-NWB T2373 experiments [<a href="#B1-aerospace-11-00213" class="html-bibr">1</a>].</p>
Full article ">
19 pages, 2390 KiB  
Article
Examination of the Influence of Alternative Fuels on Particulate Matter Properties Emitted from a Non-Proprietary Combustor
by Liam D. Smith, Joseph Harper, Eliot Durand, Andrew Crayford, Mark Johnson, Hugh Coe and Paul I. Williams
Atmosphere 2024, 15(3), 308; https://doi.org/10.3390/atmos15030308 - 29 Feb 2024
Viewed by 1098
Abstract
The aviation sector, like most other sectors, is moving towards becoming net zero. In the medium to long term, this will mean an increase in the use of sustainable aviation fuels. Research exists on the impact of fuel composition on non-volatile particulate matter [...] Read more.
The aviation sector, like most other sectors, is moving towards becoming net zero. In the medium to long term, this will mean an increase in the use of sustainable aviation fuels. Research exists on the impact of fuel composition on non-volatile particulate matter (nvPM) emissions. However, there is more sparsity when considering the impact on volatile particulate matter (vPM) emissions. Here, nine different fuels were tested using an open-source design combustor rig. An aerosol mass spectrometer (AMS) was used to examine the mass-loading and composition of vPM, with a simple linear regression algorithm used to compare relative mass spectrum similarity. The diaromatic, cycloalkane and aromatic contents of the fuels were observed to correlate with the measured total number concentration and nvPM mass concentrations, resulting in an inverse correlation with increasing hydrogen content. The impacts of fuel properties on other physical properties within the combustion process and how they might impact the particulate matter (PM) are considered for future research. Unlike previous studies, fuel had a very limited impact on the organic aerosol’s composition at the combustor exit measurement location. Using a novel combination of Positive Matrix Factorization (PMF) and high-resolution AMS analysis, new insight has been provided into the organic composition. Both the alkane organic aerosol (AlkOA) and quenched organic aerosol (QOA) factors contained CnH2n+1, CnH2n−1 and CnH2n ion series, implying alkanes and alkenes in both, and approximately 12% oxygenated species in the QOA factor. These results highlight the emerging differences in the vPM compositional data observed between combustor rigs and full engines. Full article
(This article belongs to the Section Air Pollution Control)
Show Figures

Figure 1

Figure 1
<p>Hydrogen content versus alkane, mono-aromatic and cycloalkane content of the nine fuels examined.</p>
Full article ">Figure 2
<p>Schematic of experimental setup. Note: nvPM number instrumentation and catalytic stripper described in text are not shown here for clarity as the results from these are not presented.</p>
Full article ">Figure 3
<p>Comparison of different instrument measurements (<b>A</b>)—number concentrations versus EC mass concentration, (<b>B</b>)—number concentrations versus TC and EC mass concentrations, (<b>C</b>)—EC mass concentrations) from several different aerosol instruments. Error bars show one standard deviation. The 1:1 line is shown in all graphs.</p>
Full article ">Figure 4
<p>Particle number concentrations (<b>A</b>) as a function of fuel hydrogen content, averaged across a given fuel flow rate, whilst other variables were changing (for all other variables held constant, see <a href="#app1-atmosphere-15-00308" class="html-app">Figure S5</a>). In (<b>B</b>), relative differences in hydrogen content and number concentrations are presented, the latter of which is averaged across all conditions (T1–T3 and T6–T7; excluded test points are as such due to SMPS failure for J-LA’s TP8 and J-HA’s lack of TP4–5). Values in 4b are compared against those of J-REF. X-axis error bars in 4a show one standard deviation for fuel’s repeatability of their compositional analysis. Y error bars are not shown but are SMPS measurement uncertainty (±10%). Regression values are results of linear regression with the natural log of displayed y values.</p>
Full article ">Figure 5
<p>nvPM mass concentrations (<b>A</b>) as a function of fuel hydrogen content, averaged across a given fuel flow rate, whilst other variables are changing (for all other variables held constant, see <a href="#app1-atmosphere-15-00308" class="html-app">Figure S5</a>). In (<b>B</b>), relative differences of hydrogen content and nvPM mass concentrations are presented, the latter of which is averaged across all conditions (T1–T3 and T6–T7). Both values are compared against those of J-REF. X-axis error bars in 5a show one standard deviation for fuel’s repeatability of their compositional analysis. Regression values are results of linear regression with the natural log of displayed y values.</p>
Full article ">Figure 6
<p>(<b>a</b>): Boxplot of ordinary least squares linear regression performed on organic mass concentration produced at each <span class="html-italic">m/z</span> (up to <span class="html-italic">m/z</span> 150, excluding <span class="html-italic">m/z</span> 28) by a fuel at a given condition, compared against the equivalent <span class="html-italic">m/z</span> of the organic vPM emitted by another fuel at the same operating condition, above the threshold of 0.5 µg/m<sup>3</sup>. Boxes show interquartile ranges, with median line at centre. Error bars show top and bottom quartiles of the dataset. (<b>b</b>): The same as (<b>a</b>) except <span class="html-italic">m/z</span> 12, 18, 28 and 44 were removed from comparisons.</p>
Full article ">Figure 7
<p>PMF factors and their mass concentrations as a function of MSS EC mass concentration. AlkOA (<b>left</b>) and QOA (<b>right</b>). Shaded by the hydrogen content of the fuel in the given example.</p>
Full article ">Figure 8
<p>PMF factors, categorised by their AMS ions detected via high-resolution analysis.</p>
Full article ">
22 pages, 8410 KiB  
Article
Gaps between Rice Actual and Potential Yields Based on the VPM and GAEZ Models in Heilongjiang Province, China
by Luoman Pu, Junnan Jiang, Menglu Ma and Duan Huang
Agriculture 2024, 14(2), 277; https://doi.org/10.3390/agriculture14020277 - 8 Feb 2024
Viewed by 1295
Abstract
Heilongjiang Province is a significant region for grain production and serves as a crucial commodity grain production base in China. In recent years, due to the threat of declining cropland quality and quantity, coupled with the increasingly prominent demand for grain, there is [...] Read more.
Heilongjiang Province is a significant region for grain production and serves as a crucial commodity grain production base in China. In recent years, due to the threat of declining cropland quality and quantity, coupled with the increasingly prominent demand for grain, there is an urgent need to enhance rice yields in Heilongjiang Province. It is imperative to accurately identify the gaps between actual and potential grain yields and effectively implement yield-enhancing measures in regions with significant yield gaps. This study aimed to determine the rice reproductive periods of Heilongjiang Province for 2000, 2010, and 2020, estimate the rice actual yields using the Vegetation Photosynthesis Model (VPM), simulate the rice potential yields based on the Global Agro-Ecological Zones (GAEZ) Model, and then identify the rice yield gaps at the pixel level by calculating the rice absolute yield gap (AYG) and relative yield gap (RYG). Additionally, yield-enhancing measures were proposed for regions with significant yield gaps. The results were as follows. (1) The rice reproductive periods of Heilongjiang Province for 2000, 2010, and 2020 were determined as days 153~249, days 145~249, and days 137~249. (2) The mean rice actual yield and potential yields decreased by 1222 and 5941 kg ha−1 during the 2000–2020 period, respectively, and the total actual and potential production increased by 3.75 and 1.70 million tons in Heilongjiang Province, respectively. (3) The rice AYG and RYG in the Sanjiang Plain region, such as Jixi City, Hegang City, and Jiamusi City were relatively large compared to other regions for the three years, and the rice yield gaps continued to decrease during the 2000–2020 period. (4) With regard to the Sanjiang Plain region with a large rice yield gap, this study proposes measures to narrow the rice yield gap by establishing ecological protection forests on cropland, transforming low- and middle-yielding fields, increasing agricultural science and technology inputs, selecting better rice cultivars, etc., which are important for ensuring food security. Full article
(This article belongs to the Section Crop Production)
Show Figures

Figure 1

Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>Spatial distribution of rice actual yields in 2000 based on the Vegetation Photosynthesis Model (VPM).</p>
Full article ">Figure 3
<p>Comparison between estimated rice actual yields and rice yields in the Statistical Yearbook in 2000.</p>
Full article ">Figure 4
<p>Spatial distribution of rice actual yield changes for the three periods.</p>
Full article ">Figure 5
<p>Spatial distribution of rice potential yields in 2000, 2010, and 2020 based on the Global Agro-Ecological Zones (GAEZ) model: (<b>a</b>) 2000, (<b>b</b>) 2010, and (<b>c</b>) 2020.</p>
Full article ">Figure 6
<p>Spatial distribution of rice potential yield changes in Heilongjiang Province for the three periods: (<b>a</b>) 2000–2010, (<b>b</b>) 2010–2020, and (<b>c</b>) 2000–2020.</p>
Full article ">Figure 7
<p>Correlation analysis between rice actual and potential yields for the three years: (<b>a</b>) 2000, (<b>b</b>) 2010, and (<b>c</b>) 2020.</p>
Full article ">Figure 8
<p>Absolute yield gaps (AYG) of rice actual and potential yields for the three years: (<b>a</b>) 2000, (<b>b</b>) 2010, and (<b>c</b>) 2020.</p>
Full article ">Figure 9
<p>Relative yield gaps (RYG) of rice actual and potential yields for the three years: (<b>a</b>) 2000, (<b>b</b>) 2010, and (<b>c</b>) 2020.</p>
Full article ">Figure 10
<p>Absolute yield gaps (AYG) of rice actual and potential yields in different prefecture-level cities for the three years.</p>
Full article ">Figure 11
<p>Relative yield gaps (RYG) of rice actual and potential yields in different prefecture-level cities for the three years.</p>
Full article ">
18 pages, 2663 KiB  
Article
Optimization of Temporal Coding of Tactile Information in Rat Thalamus by Locus Coeruleus Activation
by Charles Rodenkirch and Qi Wang
Biology 2024, 13(2), 79; https://doi.org/10.3390/biology13020079 - 28 Jan 2024
Viewed by 1924
Abstract
The brainstem noradrenergic nucleus, the locus coeruleus (LC), exerts heavy influences on sensory processing, perception, and cognition through its diffuse projections throughout the brain. Previous studies have demonstrated that LC activation modulates the response and feature selectivity of thalamic relay neurons. However, the [...] Read more.
The brainstem noradrenergic nucleus, the locus coeruleus (LC), exerts heavy influences on sensory processing, perception, and cognition through its diffuse projections throughout the brain. Previous studies have demonstrated that LC activation modulates the response and feature selectivity of thalamic relay neurons. However, the extent to which LC modulates the temporal coding of sensory information in the thalamus remains mostly unknown. Here, we found that LC stimulation significantly altered the temporal structure of the responses of the thalamic relay neurons to repeated whisker stimulation. A substantial portion of events (i.e., time points where the stimulus reliably evoked spikes as evidenced by dramatic elevations in the firing rate of the spike density function) were removed during LC stimulation, but many new events emerged. Interestingly, spikes within the emerged events have a higher feature selectivity, and therefore transmit more information about a tactile stimulus, than spikes within the removed events. This suggests that LC stimulation optimized the temporal coding of tactile information to improve information transmission. We further reconstructed the original whisker stimulus from a population of thalamic relay neurons’ responses and corresponding feature selectivity. As expected, we found that reconstruction from thalamic responses was more accurate using spike trains of thalamic neurons recorded during LC stimulation than without LC stimulation, functionally confirming LC optimization of the thalamic temporal code. Together, our results demonstrated that activation of the LC-NE system optimizes temporal coding of sensory stimulus in the thalamus, presumably allowing for more accurate decoding of the stimulus in the downstream brain structures. Full article
(This article belongs to the Section Neuroscience)
Show Figures

Figure 1

Figure 1
<p>Reliable response of VPM neurons to tactile stimulus. (<b>a</b>) Experimental set up. (<b>b</b>) Example raster plot and spike density function of VPm response to repeated presentation of a tactile stimulus. (<b>c</b>) Linear-nonlinear-Poisson cascade model. (<b>d</b>,<b>e</b>) Feature selectivity of spikes within and outside of events. (<b>f</b>) spikes within events transmit more information about tactile stimulus than spikes outside of events.</p>
Full article ">Figure 2
<p>LC stimulation altered the temporal structure of VPm responses. (<b>a</b>) Characterization of different types of events of an example VPm neuron. (<b>b</b>) Rate of the removed events and emerged events. (<b>c</b>) LC activation decreased the number of events. (<b>d</b>) Percentage of the removed events and emerged events. (<b>e</b>) Number of spikes per event in a trial with and without LC stimulation.</p>
Full article ">Figure 3
<p>Emerged events during LC activation transmitted more information than removed events. (<b>a</b>) Example encoded feature by spikes within removed events (orange), conserved events without LC stimulation (black), conserved events with LC stimulation (purple), and emerged events (green). (<b>b</b>) Nonlinear tuning function for the spikes within the four types of events. (<b>c</b>) Feature modulation factor for spikes within removed events vs. spikes within emerged events. (<b>d</b>) Information transmission for spikes within removed events vs. spikes within emerged events. (<b>e</b>) Feature modulation factor for spikes within conserved events without LC stimulation vs. spikes within conserved events during LC stimulation. (<b>f</b>) Information transmission for spikes within conserved events without LC stimulation vs. spikes within conserved events during LC stimulation.</p>
Full article ">Figure 4
<p>LC modulation of temporal code was not due to changes in burst firing. (<b>a</b>) Spikes within emerged events transmitted more information than spikes within removed events after bursting spikes were removed from the spike trains. (<b>b</b>) lnformation transmission per spike for all spikes, tonic spikes, bursting spikes, and burst firing occurrences without LC stimulation. (<b>c</b>) Information transmission per spike for all spikes, tonic spikes, bursting spikes, and burst firing occurrences with LC stimulation.</p>
Full article ">Figure 5
<p>The reorganization of the temporal structure of VPm events during LC activation favors ideal event placement for optimal coding of stimulus. (<b>a</b>) Example directional and non-directional feature selectivity. (<b>b</b>) Plots of feature coefficient (top) and SDF (bottom) from an example VPm neuron with directional feature selectivity illustrating ideal time points for events to occur. (<b>c</b>) Plots of feature coefficient (top) and SDF (bottom) from an example VPm neuron with non-directional feature selectivity illustrating ideal time points for events to occur. (<b>d</b>) LC activation increased the directionality of VPm feature selectivity. (<b>e</b>) LC activation increased the fraction of events that occurred at an ideal event time points.</p>
Full article ">Figure 6
<p>LC activation improved accuracy in reconstructing the original tactile stimulus from VPm population responses. (<b>a</b>) Example plot showing the original whisker stimulus, reconstructed stimulus without LC stimulation. and reconstructed stimulus with LC stimulation. (<b>b</b>) The correlation coefficient between the original and reconstructed stimulus with and without LC stimulation. (<b>c</b>) Root-mean-square deviation between the original and reconstructed stimulus.</p>
Full article ">
58 pages, 9670 KiB  
Article
Parametrization Effects of the Non-Linear Unsteady Vortex Method with Vortex Particle Method for Small Rotor Aerodynamics
by Vincent Proulx-Cabana, Guilhem Michon and Eric Laurendeau
Fluids 2024, 9(1), 24; https://doi.org/10.3390/fluids9010024 - 15 Jan 2024
Viewed by 1611
Abstract
The aim of this article is to investigate the parameter sensitivity of the (Non-Linear) Unsteady Vortex Lattice Method-Vortex Particle Method [(NL-)UVLM-VPM] with Particle Strength Exchange-Large Eddy Simulations (PSE-LES) method on a lower Reynolds number rotor. The previous work detailed the method, but introduced [...] Read more.
The aim of this article is to investigate the parameter sensitivity of the (Non-Linear) Unsteady Vortex Lattice Method-Vortex Particle Method [(NL-)UVLM-VPM] with Particle Strength Exchange-Large Eddy Simulations (PSE-LES) method on a lower Reynolds number rotor. The previous work detailed the method, but introduced parameters whose influence were not investigated. Most importantly, the Vreman model coefficient was chosen arbitrarily and was not suitable to ensure stability for this lower Reynolds number rotor simulation. In addition, the previous work presented a consistency study where geometry and time discretization were refined simultaneously. The present article starts with a comparative literature review of potential methods used to solve the aerodynamics of an isolated hovering rotor. This review highlights the differences in modeling, discretizations, sensitivity analysis, validation cases, and the results chosen by the different studies. Then, a transparent and thorough parametric study of the method is presented alongside discussions of the observed results and their physical interpretation regarding the flow. The sensitivity analysis is performed for the three free parameters of UVLM, namely Vatistas core size, the geometry and the temporal discretizations, and then for the three additional parameters introduced by UVLM-VPM, which are the Vreman model coefficient, the particle spacing, and the conversion time. The effect of different databases in the non-linear coupling is also shown. The method is shown to be consistent with both geometry and temporal refinements. It is also consistent with the expected behavior of the different parameters change, including the numerical stability that depends on the strength of the LES diffusion controlled by the Vreman model coefficient. The effect of discretization refinement presented here not only shows the integrated coefficients where different errors can cancel each other, but also looks at their convergence and where relevant, the distributed loads and tip singularity position. Finally, the aerodynamics results of the method are compared for different databases and with higher fidelity Unsteady Reynolds Averaged Navier–Stokes (URANS) 3D results on a lower Reynolds number rotor. Full article
Show Figures

Figure 1

Figure 1
<p>Time discretization problems because of the wrong core size selection.</p>
Full article ">Figure 2
<p>Visualization of the overlapping factor for the vortex particles. (<b>a</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. Each wake panel edge is replaced by one particle. The trailing particles are shown in yellow and the shed particles in purple. The arrows represent the core size selection for two different overlapping interpretations. In red is shown the core size selection of a trailing particle when the farthest neighbor is selected for the overlapping condition. In green, only the previous and next particles are selected for the overlapping condition. Only the green selection of the core size is properly adjusted to account for the new particles distance when the time step is changed at constant blade mesh.</p>
Full article ">Figure 3
<p>Visualization of the vortex particle spacing differences when the time step is changed and the panels edges are converted into a constant number of particles. (<b>a</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>, which means that each trailing particle represent an arc of 20°. (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>, so the same 20° arc has 2 particles. (<b>c</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math>, creating 4 particles over the same distance.</p>
Full article ">Figure 4
<p>Visualization of the increase of the number of particles at a coarser time step to imitate the distribution at the finest time step. (<b>a</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math> with 4 chordwise particles per iteration. (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>, with 2 chordwise particles per iteration. (<b>c</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math>, with 1 chordwise particle per iteration.</p>
Full article ">Figure 5
<p>Visualization of the tip vortex particle spacing matching <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math> at coarser time step, while reducing cluster near the root with an more even spacing. (<b>a</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math> and nTip = 4. (<b>b</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math> and nTip = 2. (<b>c</b>) <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math> and nTip = 1, which is identical to 1 chordwise particle per iteration.</p>
Full article ">Figure 6
<p>UVLM global coefficients convergence with the number of revolutions for different Vatistas core size values. The blades mesh is 8 × 20 with tip cosine refinement and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of merit (FM).</p>
Full article ">Figure 7
<p>(<b>a</b>) UVLM mean non dimensional tip singularities position with time for different Vatistas core size values and the Kocurek empirical formulation of the tip vortex position for max and min <math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math> values (<math display="inline"><semantics> <mrow> <mn>1.800</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>1.970</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math>, respectively). The top curves show radial contraction (r/R) and on the bottom the axial displacement (z/R). (<b>b</b>) the same plot with the added standard deviation of the position at every tip singularity position.</p>
Full article ">Figure 8
<p>UVLM local coefficients non dimensionnalized with section’s local velocity to better see differences in the results for the whole span of the blade for different Vatistas core size values. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 9
<p>UVLM global coefficients convergence with the number of revolutions for different blade mesh. The Vatistas core size value is 60% and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 10
<p>UVLM local coefficients for different blade mesh. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 11
<p>NL-UVLM global coefficients convergence with the number of revolutions for different time discretizations. The Vatistas core size value is 60% and blade mesh is 24 × 60. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 12
<p>NL-UVLM local coefficient of thrust evolution with the number of rotations for different time discretizations.</p>
Full article ">Figure 13
<p>NL-UVLM global coefficients convergence with the number of revolutions for different Vatistas core size values. The blade mesh is 24 × 60 and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 14
<p>NL-UVLM local coefficients for different Vatistas core size values. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 15
<p>UVLM-VPM global coefficients convergence with the number of revolutions for different Vreman model coefficients. The blade mesh is 8 × 20 and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 16
<p>UVLM-VPM global coefficients convergence with the number of revolutions for different Vreman model coefficients in a refined interval. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 17
<p>UVLM-VPM local coefficients averaged over rotations 40 to 50 for different Vreman model coefficient values. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 18
<p>UVLM-VPM local coefficients evolution averaged over rotations 40 to 50 and 60 to 70 for the smallest and largest stable Vreman model coefficient values in the refined interval. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 19
<p>UVLM-VPM mean non dimensional tip singularities position with time averaged between rotations 40 to 50 for different Vreman model coefficient values. The panel from <a href="#sec5dot1-fluids-09-00024" class="html-sec">Section 5.1</a> result is shown as reference.</p>
Full article ">Figure 20
<p>NL-UVLM-VPM global coefficients convergence with the number of revolutions for different tip vortex particle spacing. The Vreman model coefficient is set according to the relation found in the previous section. The blade mesh is 24 × 60 and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 21
<p>NL-UVLM-VPM global coefficients convergence with the number of revolutions for different tip vortex particle spacing. The blade mesh is 24 × 60 and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 22
<p>NL-UVLM-VPM local coefficient of thrust evolution for different averaging windows. <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 23
<p>NL-UVLM-VPM local coefficients for different tip vortex particle spacing compared with the panel reference. <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 24
<p>NL-UVLM-VPM mean non dimensional tip singularities position with time for different tip vortex particle spacing. The panel from <a href="#sec5dot1-fluids-09-00024" class="html-sec">Section 5.1</a> result is shown as reference.</p>
Full article ">Figure 25
<p>NL-UVLM-VPM global coefficients convergence with the number of revolutions for different blade mesh. <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>20</mn> <mo>°</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 26
<p>NL-UVLM-VPM local coefficients for different blade mesh. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 27
<p>NL-UVLM-VPM global coefficients convergence with the number of revolutions for different time discretizations. The blade mesh is 24 × 60 and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 28
<p>NL-UVLM-VPM local coefficients for different time discretizations. <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>5</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 29
<p>NL-UVLM-VPM global coefficients convergence with the number of revolutions for different numbers of wake panels rotations before being transformed in particles. The blade mesh is 24 × 60, <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mo>Ψ</mo> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>x</mi> <mrow> <mi>t</mi> <mi>i</mi> <mi>p</mi> </mrow> </msub> <mo>=</mo> <mn>10</mn> <mo>°</mo> </mrow> </semantics></math>. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 30
<p>NL-UVLM-VPM local coefficients for different numbers of wake panels rotations before being transformed in particles. (<b>a</b>) Local coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>t</mi> </msub> </semantics></math>). (<b>b</b>) Local coefficient of torque (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>q</mi> </msub> </semantics></math>).</p>
Full article ">Figure 31
<p>NL-UVLM-VPMmean non dimensional tip singularities position with time for different numbers of wake panels rotations before being transformed in particles. The panel from <a href="#sec5dot1-fluids-09-00024" class="html-sec">Section 5.1</a> result is shown as reference.</p>
Full article ">Figure 32
<p>Comparison of the convergence of the global coefficients with the number of revolutions between the different viscous database and the 3D URANS reference. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure 33
<p>Thrust coefficient distribution for the different NL-UVLM-VPM databases and the URANS reference.</p>
Full article ">Figure 34
<p>Torque coefficient distribution for the different NL-UVLM-VPM databases and the URANS reference.</p>
Full article ">Figure 35
<p>Twisting moment about the half chord distribution for the different NL-UVLM-VPM databases and the URANS reference.</p>
Full article ">Figure A1
<p>Coefficient. of lift for the profiles at the two ends of the blade computed with different software and turbulence model. (<b>a</b>) Root section at Re = 26,884. (<b>b</b>) Tip section at Re = 170,263.</p>
Full article ">Figure A2
<p>Coefficient of drag for the profiles at the two ends of the blade computed with different software and turbulence model. (<b>a</b>) Root section at Re = 26,884. (<b>b</b>) Tip section at Re = 170,263.</p>
Full article ">Figure A3
<p>Coefficient of moment for the profiles at the two ends of the blade computed with different software and turbulence model. (<b>a</b>) Root section at Re = 26,884. (<b>b</b>) Tip section at Re = 170,263.</p>
Full article ">Figure A4
<p>Stream traces of the NSCODE SA simulation at the root Reynolds number (Re = 26,884) and 2 degrees of angle of attack showing a separation bubble on the upper and lower sides.</p>
Full article ">Figure A5
<p>3D URANS convergence of the global coefficients with the number of revolutions for the different meshes. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure A6
<p>3D URANS convergence of the global coefficients with the number of revolutions for the different time discretizations. (<b>a</b>) Coefficient of thrust (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>T</mi> </msub> </semantics></math>). (<b>b</b>) Figure of Merit (FM).</p>
Full article ">Figure A7
<p>3D URANS mean thrust coefficient distribution obtained with the different meshes.</p>
Full article ">Figure A8
<p>3D URANS mean torque coefficient distribution obtained with the different meshes.</p>
Full article ">Figure A9
<p>3D URANS mean twisting moment about half chord distribution obtained with the different meshes.</p>
Full article ">
25 pages, 6033 KiB  
Article
Improved Modeling of Gross Primary Production and Transpiration of Sugarcane Plantations with Time-Series Landsat and Sentinel-2 Images
by Jorge Celis, Xiangming Xiao, Paul M. White, Osvaldo M. R. Cabral and Helber C. Freitas
Remote Sens. 2024, 16(1), 46; https://doi.org/10.3390/rs16010046 - 21 Dec 2023
Cited by 1 | Viewed by 1363
Abstract
Sugarcane croplands account for ~70% of global sugar production and ~60% of global ethanol production. Monitoring and predicting gross primary production (GPP) and transpiration (T) in these fields is crucial to improve crop yield estimation and management. While moderate-spatial-resolution (MSR, hundreds of meters) [...] Read more.
Sugarcane croplands account for ~70% of global sugar production and ~60% of global ethanol production. Monitoring and predicting gross primary production (GPP) and transpiration (T) in these fields is crucial to improve crop yield estimation and management. While moderate-spatial-resolution (MSR, hundreds of meters) satellite images have been employed in several models to estimate GPP and T, the potential of high-spatial-resolution (HSR, tens of meters) imagery has been considered in only a few publications, and it is underexplored in sugarcane fields. Our study evaluated the efficacy of MSR and HSR satellite images in predicting daily GPP and T for sugarcane plantations at two sites equipped with eddy flux towers: Louisiana, USA (subtropical climate) and Sao Paulo, Brazil (tropical climate). We employed the Vegetation Photosynthesis Model (VPM) and Vegetation Transpiration Model (VTM) with C4 photosynthesis pathway, integrating vegetation index data derived from satellite images and on-ground weather data, to calculate daily GPP and T. The seasonal dynamics of vegetation indices from both MSR images (MODIS sensor, 500 m) and HSR images (Landsat, 30 m; Sentinel-2, 10 m) tracked well with the GPP seasonality from the EC flux towers. The enhanced vegetation index (EVI) from the HSR images had a stronger correlation with the tower-based GPP. Our findings underscored the potential of HSR imagery for estimating GPP and T in smaller sugarcane plantations. Full article
(This article belongs to the Special Issue Remote Sensing of Primary Production)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The two sugarcane locations with EC flux tower sites (red doted polygons; (<b>a</b>) USA and (<b>b</b>) Brazil) displaying the pixel size of the optical data utilized in the study: 10 m (red polygon, Sentinel-2), 30 m (blue polygon, Landsat), and 500 m (green polygon, MODIS).</p>
Full article ">Figure 2
<p>Seasonal variation in daily temperature, daily mean daytime temperature, 8-day average photosynthetically active radiation (PAR), and accumulated rainfall (8-day interval) of the sugarcane EC flux tower sites. (<b>a</b>) FAYS-Brazil site 2015–2017. (<b>b</b>) Chacahoula, USA site 2018–2020.</p>
Full article ">Figure 3
<p>Seasonal variation in estimated GPP<sub>EC</sub> and measured NEE<sub>EC</sub> at 8-day intervals over the study period. Seasonal variation in vegetation indices (land surface water index (LSWI) and enhanced vegetation index (EVI)) derived from 8-day MODIS data. (<b>a</b>,<b>c</b>) FAYS Brazil site, 2015–2017. (<b>b</b>,<b>d</b>) Chacahoula, USA site (2018–2020).</p>
Full article ">Figure 4
<p>(<b>Left panel</b>): The relationships between gross primary production (GPP<sub>EC</sub>) and the enhanced vegetation index (EVI) derived from moderate-spatial-resolution images (MSR, MODIS) and high-spatial-resolution images (HSR, Landsat and Sentinel-2) within the growing season. (<b>Right panel</b>): 3 m 4-band planet Scope and 5 m RapidEye Ortho tile surface reflectance true-color images of vegetation cover in the FAYS Brazil study area. The red circle represents the area used to obtain the time-series of HSR data.</p>
Full article ">Figure 5
<p>(<b>Left panel</b>): The relationship between the estimated gross primary production (GPP<sub>EC</sub>) and enhanced vegetation index derived from moderate-spatial-resolution images (MSR, MODIS) and high-spatial-resolution images (HSR, Landsat and Sentinel-2) within the growing season. (<b>Right panel</b>): 3 m 4-band planet Scope and 5 m RapidEye Ortho tile surface reflectance true-color images of vegetation cover in the Chacahoula, USA study area. The red circle represents the area used to obtain the time-series of HSR data.</p>
Full article ">Figure 6
<p>Relationships between estimated gross primary production (GPP_EC), mean daily temperature, and mean daily daytime temperature within the sugarcane growing seasons. (<b>a</b>,<b>c</b>) FAYS Brazil site, 2015–2017. (<b>b</b>,<b>d</b>) Chacahoula, USA site (2018–2020).</p>
Full article ">Figure 7
<p>Relationship between enhanced vegetation index derived from MODIS (EVI 500 m) and Landsat/Sentinel-2 (EVI 10 m) with mean daily air temperature and mean daily daytime temperature during the growing seasons. (<b>a</b>,<b>b</b>) FAYS Brazil site, 2015–2017. (<b>c</b>,<b>d</b>) Chacahoula, USA site (2018–2020).</p>
Full article ">Figure 8
<p>GPP estimate time series from the EC flux sites (black line) and the VPM (HSR GPP in blue and MSR GPP in red). (<b>a</b>) FAYS Brazil site GPP time series estimates (2015–2017). (<b>b</b>) Chacahoula, USA site GPP time series estimates (2018–2020). Relationship between GPP<sub>EC</sub> and the VPM (HSR GPP in blue and MSR GPP in red) (<b>c</b>) FAYS Brazil. (<b>d</b>) Chacahoula, USA site.</p>
Full article ">Figure 9
<p>ET<sub>EC</sub> estimate time series from the EC flux sites (black line) and the T<sub>VTM_EC</sub> (red line) for (<b>a</b>) the FAYS Brazil site and (<b>b</b>) Chacahoula, USA.</p>
Full article ">
17 pages, 3598 KiB  
Article
Batch Settling and Low-Pressure Consolidation Behaviors of Dredged Mud Slurry: Steady-State Evaluation and Mechanism Study
by Shufeng Bao, Lingfeng Guo, Zhiliang Dong, Ruibo Zhou, Shuangxi Zhou and Jian Chen
Water 2024, 16(1), 7; https://doi.org/10.3390/w16010007 - 19 Dec 2023
Viewed by 1188
Abstract
Since the exploration of the characteristics of dredged mud slurry during batch settlement and low-pressure consolidation (less than 100 kPa) is still insufficient, the determination of the optimal time to start the vacuum preloading method (VPM) on dredged-fill foundations is still empirically oriented [...] Read more.
Since the exploration of the characteristics of dredged mud slurry during batch settlement and low-pressure consolidation (less than 100 kPa) is still insufficient, the determination of the optimal time to start the vacuum preloading method (VPM) on dredged-fill foundations is still empirically oriented (due to a lack of enough scientific basis). To further explore the characteristics of dredged mud slurry during batch settlement and low-pressure consolidation, samples from typical dredged-fill land projects were obtained and used to conduct batch sedimentation model experiments and low-pressure (less than 100 kPa) consolidation tests. The results of experiments and analyses showed the following: (1) the clay (d < 0.005 mm) content is a main factor affecting the batch settlement and consolidation characteristics of dredged mud slurry, which is not conducive to the consolidation effect of dredged-fill foundations. (2) For dredged mud slurry whose clay content is within 40% to 60%, the cumulative change rate of the average porosity ratio of 60% to 75% is suitable for evaluating the steady state of its batch sedimentation process, i.e., the optimal starting time of VPM. Finally, based on the experimental analyses, a settlement prediction method that considers both the batch sedimentation and the low-pressure consolidation processes was developed and validated. Full article
Show Figures

Figure 1

Figure 1
<p>A photo showing the filling process of dredged soil during a land reclamation project.</p>
Full article ">Figure 2
<p>A map showing the sample sources for the experiments. The map is originally from <a href="https://www.amap.com/" target="_blank">https://www.amap.com/</a>, accessed on 10 December 2023.</p>
Full article ">Figure 3
<p>Schematic diagram of the experiment apparatus.</p>
Full article ">Figure 4
<p>Schematic diagram of the static batch settling process.</p>
Full article ">Figure 5
<p>Loading schedule of the low-pressure consolidation test. The loading time at all levels also refers to the time required for the sample with a height of 2 cm in Standard for Geotechnical Testing Method GB/T 50123-2019 [<a href="#B30-water-16-00007" class="html-bibr">30</a>]. Among them, the loading stage of 1 to 11 kPa was set to prevent the soil sample from being crushed.</p>
Full article ">Figure 6
<p>The settlement curve of the solid–liquid interface <span class="html-italic">Si</span> over time.</p>
Full article ">Figure 7
<p>The void ratio variation curves of the low-pressure consolidation tests. (<b>a</b>) <span class="html-italic">e</span>-<span class="html-italic">P</span> curves; (<b>b</b>) <span class="html-italic">e</span>-log<span class="html-italic">P</span> curves.</p>
Full article ">Figure 8
<p>The physical and mechanical test results of the low-pressure consolidation tests. (<b>a</b>) Water content; (<b>b</b>) Density; (<b>c</b>) Undrained shear strength; (<b>d</b>) 3D display.</p>
Full article ">Figure 9
<p>The schematic graphs of the entire batch sedimentation process. (<b>a</b>) Settling process of mixed particle groups with different particle sizes and density; (<b>b</b>) The flow field between two soil particles during the hindered settling.</p>
Full article ">Figure 10
<p>Changes in particle flux during intermittent settling.</p>
Full article ">Figure 11
<p>The curves of the cumulative change rate <span class="html-italic">R<sub>e</sub></span> of the average pore ratio of samples from different regions. The water content and clay content of the referenced dredged soil from Binhai, Tianjin [<a href="#B39-water-16-00007" class="html-bibr">39</a>] are respectively 400% and 47%.</p>
Full article ">
34 pages, 32930 KiB  
Article
Thalamic Neuron Resilience during Osmotic Demyelination Syndrome (ODS) Is Revealed by Primary Cilium Outgrowth and ADP-ribosylation factor-like protein 13B Labeling in Axon Initial Segment
by Jacques Gilloteaux, Kathleen De Swert, Valérie Suain and Charles Nicaise
Int. J. Mol. Sci. 2023, 24(22), 16448; https://doi.org/10.3390/ijms242216448 - 17 Nov 2023
Viewed by 1498
Abstract
A murine osmotic demyelinating syndrome (ODS) model was developed through chronic hyponatremia, induced by desmopressin subcutaneous implants, followed by precipitous sodium restoration. The thalamic ventral posterolateral (VPL) and ventral posteromedial (VPM) relay nuclei were the most demyelinated regions where neuroglial damage could be [...] Read more.
A murine osmotic demyelinating syndrome (ODS) model was developed through chronic hyponatremia, induced by desmopressin subcutaneous implants, followed by precipitous sodium restoration. The thalamic ventral posterolateral (VPL) and ventral posteromedial (VPM) relay nuclei were the most demyelinated regions where neuroglial damage could be evidenced without immune response. This report showed that following chronic hyponatremia, 12 h and 48 h time lapses after rebalancing osmolarity, amid the ODS-degraded outskirts, some resilient neuronal cell bodies built up primary cilium and axon hillock regions that extended into axon initial segments (AIS) where ADP-ribosylation factor-like protein 13B (ARL13B)-immunolabeled rod-like shape content was revealed. These AIS-labeled shaft lengths appeared proportional with the distance of neuronal cell bodies away from the ODS damaged epicenter and time lapses after correction of hyponatremia. Fine structure examination verified these neuron abundant transcriptions and translation regions marked by the ARL13B labeling associated with cell neurotubules and their complex cytoskeletal macromolecular architecture. This necessitated energetic transport to organize and restore those AIS away from the damaged ODS core demyelinated zone in the murine model. These labeled structures could substantiate how thalamic neuron resilience occurred as possible steps of a healing course out of ODS. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental groups of mice undergoing chronic hyponatremia and correction of hyponatremia.</p>
Full article ">Figure 2
<p>Murine brain parasagittal paraffin-sections stained with Eriochrome C (<b>left</b>) and immunostained against MBP protein (<b>right</b>). The section of the ODS48h brain where the thalamus (th) clearly reveals a poor contrast caused by the extrapontine regional demyelination. Scale bar represents 1 mm.</p>
Full article ">Figure 3
<p>LM paraffin section of ODS12h ARL13b immunolabeled thalamus aspect. The upper left broken line delimitates a main part of the ODS damaged zone (star). Adjacent to and more distally, the thalamus field of view reveals many examples of some shafts or rod-shaped labeled axon initial segments (AIS), marked by white arrows among other cells and structures; As: Astrocyte; c: capillary; N: neuron; O: oligodendrocyte. Insert: an AIS can be recognized associating with joined thalamic neurons. Scales = 10 µm.</p>
Full article ">Figure 4
<p>ARL13B immunolabeling of untreated (NN), chronic hyponatremic (HN,) and ODS murine thalamus VPL 12 h (ODS12h a and b) and 48 h (ODS48h a and b) post-treatment with hemalum counterstain. White arrows mark some examples of AIS, shaped from specks to rod-like straight or curved appendices in each frame of the pane that is associated with HN, ODS12h, and ODS48h thalamic nerve cell bodies (N) As: astrocyte; black arrows indicate some of the oligodendrocytes; black star indicates rod-like appendice emerging from oligodendrocyte; c: capillary. All the scales = 10 µm.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>): Samples of counts and measurements of ARL13B AIS labeled from 12,500 µm<sup>2</sup> LM fields of view compared between NN, HN, ODS12h, and ODS48h treatments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, ns = not significant.</p>
Full article ">Figure 6
<p>(<b>A</b>–<b>C</b>): LM aspects of 12 h ODS murine thalamus within 100 µm distance from the ODS damaged rim. (<b>A</b>): ARL13b immunolabeled depicting a small rod-shaped appendix (white arrow) to a nerve cell body (N) where hemalum has enhanced RNA in pale blue more than the DNA content of nucleolus. (<b>B</b>,<b>C</b>): One-µm thick epoxy section views of nerve cell bodies (N) where similar heavily basophilic shapes appeared issued from the narrow perikaryal zone; m: myelinated nerve bundles with diverse orientations. All the scales = 10 µm.</p>
Full article ">Figure 7
<p>(<b>A</b>–<b>C</b>): <b>A</b>: TEM aspect of 12hODS murine thalamus within 100 µm distance from the ODS’ damaged edge revealing a neuron with a fine structure like those of <a href="#ijms-24-16448-f006" class="html-fig">Figure 6</a>A–C with perikaryal content of an encircling Golgi apparatus (G) and especially containing a zone enriched with granulated content associated with membranes as RER (as revealed by (<b>B</b>)). The apex view region indicated by a white ellipsoid in A and marked by the white arrow had a perikaryon contrasted with granulations. As shown in an enlarged view of <b>C</b>, a neuroplasm congested by free and polyribonucleoproteins attached to endoplasmic reticulum; scale equals 25 nm. Note the stars adjacent to nuclear envelope pores marked adjacent neuroplasm damages. The nucleolus with numerous nucleolar organizer centers (NORs) made of dense and fine fibrils (f) and the large mass of RNA storage associated as granular center (g).</p>
Full article ">Figure 8
<p>(<b>A</b>–<b>E</b>): (<b>A</b>): Parasagittal LM paraffin section part of ODS48h thalamic VPL region as viewed following blood–brain barrier immunoglobulin G permeability assessement [<a href="#B67-ijms-24-16448" class="html-bibr">67</a>], with the epicenter of the demyelinated region outlined; scale is 500 µm. (<b>B</b>): One-µm thick epoxy section, toluidine blue-stained of an area of the ODS demyelinated zone, shown filled with diverse sized vacuoles resulting from tissue degradations. M: microglial cell, N: neuron, O: oligodendrocyte; scale = 20 µm. (<b>C</b>–<b>E</b>): TEM aspects of similar spoiled neuropil (Np) ODS zone as in (<b>B</b>) where (<b>C</b>) showed damaged neurons (N) with diluted nucleoplasm and vacuoles (star) and lysosomes in perikaryal zones that included a necrotic oligodendrocyte satellite (O). (<b>D</b>): a nerve cell body with axon hillock, scale = 5 µm. (<b>E</b>): TEM enlarged aspect of (<b>D</b>) where the axon hillock (Ah) transition to the initial segment (AIS) is marked by a white ellipse, revealing microtubules among mitochondria (mt), and other associated organelles that became aligned in the latter, also underlined by neuroplasm gaps among the cytoskeleton (white stars). White arrow points a part of the AIS and black arrows indicate the local enrichment in neurotubules.</p>
Full article ">Figure 9
<p>The enlarged view of the AIS segment, part in fine structure, shown in <a href="#ijms-24-16448-f008" class="html-fig">Figure 8</a>E that illustrates the rich proteinaceous content of the thalamic cell body extension where the apparent blur-like view showed a rich microtubule core positioned into paraxial and parallel alignment and a tangent section profile of mitochondrion (mt) was shown and some cytoskeletal cross-linkages divulged their periodical architecture (small white arrows and arrowhead sets); in the profile, the sub neurolemma was underlined by a near distant delicate, neuroplasm protein arrangement (insert and white arrowheads); white open arrows indicate the swirled or twisted endoplasmic reticulum parts within the AIS crowded cytoskeleton. High contrast granules, 24–27 nm diameter, should be ribonucleoprotein strings.</p>
Full article ">Figure 10
<p>(<b>A</b>–<b>C</b>): TEM of an example of ODS48h resilient thalamic neuron (N), located in the ODS damaged zone, demonstrating a primary cilium (white arrows) among the degraded neuropil (<b>A</b>), and in its enlarged aspect (<b>B</b>) showing the narrow neck pocket (black arrow) whose basal aspect contains arrays of fibrillar structures that reach a hub-like area where a fascicle of microtubules also known as neurotubules issue from the perikaryon attained; these and other organelles (mt: mitochondrion) were shrouded by numerous ribonucleoproteins. Additionally, in both (<b>A</b>,<b>B</b>) illustrations, black stars mark diverse delicate lamellipodia, adjacent to the primary cilium. (<b>C</b>): filopodia (arrows) noticed adjacent to the primary cilium.</p>
Full article ">Figure 11
<p>(<b>A</b>–<b>D</b>): TEM aspects of ODS48h thalamic nerve cell bodies located outside the outskirts of demyelination. A: Neuron (N) with its axon hillock (Ah) and twisted AIS (white arrow) among a neuropil where scattered, small intercellular spaces remain from ODS damage (stars) and a satellite oligodendrocyte (O). (<b>B</b>–<b>D</b>): Axon hillock regions reveal Golgi apparatus (G), mitochondria with winding endoplasm among free and attached polyribosomes. In (<b>B</b>), the AIS part with synaptic contact (white arrowhead) and adjacent astrocyte foot (As) are shown.</p>
Full article ">Figure 12
<p>(<b>A</b>,<b>B</b>): TEM aspects of ODS48h thalamic neurons. (<b>A</b>): AIS early segment issued from axon hillock (Ah) marked by white arrows indicating synaptic connections. (<b>B</b>): Enlarged part of transition of axon hillock to AIS segment recognized by Golgi saccules (G) erupting with numerous vesicles and adjacent neurotubules that resolve from erratic orientation to a more paraxial to parallel funnel shape fascicle (white arrowheads), where vesicles (black arrowheads) located adjacent to neurolemma as cisternal organelles and along mitochondria (mt), typically without myelin, are surrounded by a few end-feet of astrocytes (As), recognized by glycogen particle contents.</p>
Full article ">Figure 13
<p>(<b>A</b>–<b>C</b>): Thalamic ODS48h nerve cell bodies and axon initial segment (IS). (<b>A</b>): LM view where two joined neurons have one on the right that displays an ARL13B labeled rod-like AIS structure (white arrow). N: neuron; O: oligodendrocyte. A bracket encompasses the TEM corresponding aspect in (<b>B</b>) with further sectioning level and enlargement in (<b>C</b>) if the axon IS among the neuropils, still depicting ODS damage remnants (stars); black arrowheads indicate astrocyte parts and white arrowheads mark synaptic contacts.</p>
Full article ">Figure 14
<p>(<b>A</b>–<b>F</b>): Thalamic ODS48h nerve cell bodies and elongated axon initial segment (white arrows) aspects viewed with LM (<b>A</b>,<b>B</b>) and TEM (<b>C</b>–<b>F</b>). (<b>A</b>): ARL13B labeled AIS shown of two neurons as shaft and curved rod and in (<b>B</b>), from 1-µm thick epoxy section, toluidine blue-stained. (<b>C</b>,<b>D</b>): AIS from resilient neurons (N), including insert in (<b>D</b>), to show neurotubules’ bundle; O: oligodendrocyte. (<b>E</b>,<b>F</b>): Elongated AIS among the neuropils, containing huge mitochondria profiles and some synaptic contacts can be seen (white arrowheads).</p>
Full article ">Figure 15
<p>Western blots combining samples of NN, HN, and ODS48h thalamus where ARL13B protein is detected across all samples without significant variation; GAPDH is included as a loading control. A protein sample from mouse testis was used as positive (pos.) control.</p>
Full article ">Figure A1
<p>Validation of ARL13B immunolabeling specificity. Testis seminiferous tubule flagella (<b>A</b>–<b>C</b>), and the stereocilia in proximal segment of epididymis (<b>D</b>,<b>E</b>) of the same murine used for brain ODS experiments. The ARL13B immunolabeling associated with these appendages and their subcellular maintenance to reveal labels in (<b>B</b>,<b>C</b>,<b>E</b>), counterstained with hemalum. Scales = 10 µm.</p>
Full article ">Figure A2
<p>TEM of ODS48h thalamic neuron revealed euchromatic nucleus with indent (<b>A</b>) where an activate nucleolus revealed in an enlarged view of (<b>B</b>) the numerous transcripts poured out (arrows) through nuclear pores as ribonucleoproteins seen in the perikaryal indent (star).</p>
Full article ">Figure A3
<p>Western blots of ARL13B (<b>A</b>) and GAPDH (<b>B</b>) protein levels of the thalamus tissue samples as absorbed in laboratory preparations.</p>
Full article ">
17 pages, 13946 KiB  
Article
The Spatio-Temporal Variations of GPP and Its Climatic Driving Factors in the Yangtze River Basin during 2000–2018
by Chong Nie, Xingan Chen, Rui Xu, Yanzhong Zhu, Chenning Deng and Queping Yang
Forests 2023, 14(9), 1898; https://doi.org/10.3390/f14091898 - 18 Sep 2023
Cited by 1 | Viewed by 1103
Abstract
Terrestrial gross primary productivity (GPP) is the major carbon input to the terrestrial ecosystem. The Yangtze River Basin (YRB) holds a key role in shaping China’s economic and social progress, as well as in ecological and environmental protection. However, how the GPP in [...] Read more.
Terrestrial gross primary productivity (GPP) is the major carbon input to the terrestrial ecosystem. The Yangtze River Basin (YRB) holds a key role in shaping China’s economic and social progress, as well as in ecological and environmental protection. However, how the GPP in the YRB responds to the climate factors remain unclear. In this research, we applied the Vegetation Photosynthesis Model (VPM) GPP data to explore the spatial and temporal variations of GPP in the YRB during 2000–2018. Based on the China Meteorological Forcing Dataset (CMFD), the partial least squares regression (PLSR) method was employed to identify the GPP responses to changes in precipitation, temperature, and shortwave radiation between 2000 and 2018. The findings showed that the long-term average of GPP in the YRB was 1153.5 ± 472.4 g C m−2 yr−1 between 2000 and 2018. The GPP of the Han River Basin, the Yibin-Yichang section of the Yangtze River mainstream, and the Poyang Lake Basin were relatively high, while the GPP of the Jinsha River Basin above Shigu and the Taihu Lake Basin were relatively low. A significant upward trend in GPP was observed over the 19-year period, with an annual increase rate of 8.86 g C m−2 yr−1 per year. The GPP of the Poyang Lake Basin and Jialing River Basin grew much faster than other water resource regions. Savannas and forests also had relatively higher GPP rate of increase compared to other vegetation types. The relative contributions of precipitation, temperature, and shortwave radiation to GPP variations in the YRB were 13.85 ± 13.86%, 58.87 ± 9.79%, and 27.07 ± 15.92%, respectively. Our results indicated that temperature was the main climatic driver on the changes of GPP in the YRB. This study contributes to an in-depth understanding of the variations and climate-impacting factors of vegetation productivity in the YRB. Full article
(This article belongs to the Section Forest Meteorology and Climate Change)
Show Figures

Figure 1

Figure 1
<p>The location and 12 water resource regions of the YRB (<b>a</b>). The spatial distributions of long-term averages of precipitation (<b>b</b>), temperature (<b>c</b>), shortwave radiation (<b>d</b>), and vegetation cover (<b>e</b>) of the YRB.</p>
Full article ">Figure 1 Cont.
<p>The location and 12 water resource regions of the YRB (<b>a</b>). The spatial distributions of long-term averages of precipitation (<b>b</b>), temperature (<b>c</b>), shortwave radiation (<b>d</b>), and vegetation cover (<b>e</b>) of the YRB.</p>
Full article ">Figure 2
<p>The spatial variations of annual average GPP in the YRB between 2000 and 2018 (<b>a</b>) and its statistical characteristics (<b>b</b>). The variations of the long-term average of GPP with the changes in latitude (<b>c</b>) and longitude (<b>d</b>).</p>
Full article ">Figure 3
<p>Boxplot of the variations of GPP in different water resource regions (<b>a</b>), and vegetation types (<b>b</b>) in the YRB. Boxplot elements: square points = mean values; horizontal lines within boxes = medians; whiskers = 1.5 times of interquartile ranges; and points outside the boxes = outliers.</p>
Full article ">Figure 4
<p>The variations of annual average GPP (<b>a</b>), precipitation (<b>b</b>), temperature (<b>c</b>), and shortwave radiation (<b>d</b>) from 2000 to 2018 in the YRB.</p>
Full article ">Figure 5
<p>The variations of GPP change rate (<b>a</b>), the significance of regression analysis (<b>b</b>), the boxplot of the GPP change rate in 12 water resource regions (<b>c</b>), and the boxplot of GPP change rate in different vegetation types (<b>d</b>). “Increase*” represented that the increasing trends were significant (<span class="html-italic">p</span> &lt; 0.1). “Increase” represented that the increasing trends were insignificant (<span class="html-italic">p</span> &gt; 0.1). “Decrease*” represented that the decreasing trends were significant (<span class="html-italic">p</span> &lt; 0.1). “Decrease” represented that the decreasing trends were insignificant (<span class="html-italic">p</span> &gt; 0.1). Boxplot elements: square points = mean values; horizontal lines within boxes = medians; whiskers = 1.5 times of interquartile ranges; and points outside the boxes = outliers. Purple lines represented the “zero value” lines.</p>
Full article ">Figure 5 Cont.
<p>The variations of GPP change rate (<b>a</b>), the significance of regression analysis (<b>b</b>), the boxplot of the GPP change rate in 12 water resource regions (<b>c</b>), and the boxplot of GPP change rate in different vegetation types (<b>d</b>). “Increase*” represented that the increasing trends were significant (<span class="html-italic">p</span> &lt; 0.1). “Increase” represented that the increasing trends were insignificant (<span class="html-italic">p</span> &gt; 0.1). “Decrease*” represented that the decreasing trends were significant (<span class="html-italic">p</span> &lt; 0.1). “Decrease” represented that the decreasing trends were insignificant (<span class="html-italic">p</span> &gt; 0.1). Boxplot elements: square points = mean values; horizontal lines within boxes = medians; whiskers = 1.5 times of interquartile ranges; and points outside the boxes = outliers. Purple lines represented the “zero value” lines.</p>
Full article ">Figure 6
<p>The spatial distributions of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>) in the YRB. Boxplots of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> in 12 water resource regions (<b>d</b>) and vegetation types (<b>e</b>). Boxplot elements: solid points = mean values; horizontal lines within boxes = medians; whiskers = 1.5 times of interquartile ranges; and “+” outside the boxes = outliers. Blue lines represented the “zero value” lines.</p>
Full article ">Figure 7
<p>The spatial distributions of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>) in the YRB. Histograms of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> in 12 water resource regions (<b>d</b>) and vegetation types (<b>e</b>).</p>
Full article ">Figure 7 Cont.
<p>The spatial distributions of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> (<b>a</b>), <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math> (<b>b</b>) and <math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> (<b>c</b>) in the YRB. Histograms of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ξ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> in 12 water resource regions (<b>d</b>) and vegetation types (<b>e</b>).</p>
Full article ">
Back to TopTop