[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (258)

Search Parameters:
Keywords = SST coefficients

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 2779 KiB  
Article
Wind Turbine Enhancement via Active Flow Control Implementation
by Marc Lahoz, Ahmad Nabhani, Mohammad Saemian and Josep M. Bergada
Appl. Sci. 2024, 14(23), 11404; https://doi.org/10.3390/app142311404 - 7 Dec 2024
Viewed by 531
Abstract
The present research enhances the efficiency of an airfoil section from the DTU-10MW Horizontal Axis Wind Turbine (HAWT) via Active Flow Control (AFC) implementation and when using synthetic jets (SJ). The flow around two airfoil sections cut along the wind turbine blade and [...] Read more.
The present research enhances the efficiency of an airfoil section from the DTU-10MW Horizontal Axis Wind Turbine (HAWT) via Active Flow Control (AFC) implementation and when using synthetic jets (SJ). The flow around two airfoil sections cut along the wind turbine blade and for a wind speed of 10 m/s is initially simulated using the CFD-2D-RANS-Kω-SST turbulence model, from where the time-averaged boundary layer separation point and the associated vortex shedding frequency are obtained. On a second stage of the paper, and considering one of the two airfoil sections, the boundary layer separation point previously determined is used to locate the SJ groove as well as the groove width; the three remaining AFC parameters, momentum coefficient, jet inclination angle, and jet pulsating frequency, are parametrically optimized. Thanks to the energy assessment presented in the final part of the paper, the study shows that a considerable power increase of the airfoil section can be obtained when attaching the former separated boundary layer. The extension of the optimization process to the rest of the blade sections where the boundary layer is separated would lead to an efficiency increase of the HAWT. The Reynolds numbers associated to the respective airfoil sections analyzed in the present manuscript are Re = 14.088×106 and Re = 14.877×106, the characteristic length being the corresponding chord length for each airfoil. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic view of the meshing domain within its general dimensions.</p>
Full article ">Figure 2
<p>Three-dimensional view of the DTU10MW blade with the two sections studied.</p>
Full article ">Figure 3
<p>Blade diagram with the different airfoil profiles along the blade span.</p>
Full article ">Figure 4
<p>Full view of the mesh in the computational domain (<b>a</b>), zoomed view of the structured mesh in the leading ((<b>b</b>) <b>top</b>) and trailing ((<b>b</b>) <b>bottom</b>) edges for FFA-W3-301 airfoil type.</p>
Full article ">Figure 5
<p>Synthetic jet position and associated parameters (<b>a</b>). Detailed view of the AFC mesh implementation in Section 54 (<b>b</b>).</p>
Full article ">Figure 6
<p>Pressure (<b>a</b>) and friction (<b>b</b>) coefficients comparison for the four meshes evaluated in Section 54.</p>
Full article ">Figure 7
<p>Wind, rotor velocities, and forces on a generic blade section.</p>
Full article ">Figure 8
<p>Velocity fields given as time-averaged streamlines and time-averaged pressure fields for Sections 54 and 81.</p>
Full article ">Figure 9
<p>Post-processing results for Sections 54 and 81. Pressure coefficient distribution along upper and lower airfoil surfaces (<b>a</b>). Tangential friction coefficient distribution along airfoils upper surfaces (<b>b</b>). Mean velocity distribution tangent to the airfoils upper surfaces and as a function of the normalized wall normal distance (<b>c</b>). Fast Fourier Transform (FFT) of lift coefficients (<b>d</b>).</p>
Full article ">Figure 10
<p>Lift coefficient (<b>a</b>), drag coefficient (<b>b</b>), and aerodynamic efficiency (<b>c</b>) variation as a function of momentum coefficient; the vertical bars characterize the corresponding peak-to-peak amplitude. Constant jet angle (<math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>jet</mi> </msub> <mo>=</mo> <mn>10</mn> <mspace width="0.222222em"/> </mrow> </semantics></math> deg.) and forcing frequency (<math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mo>+</mo> </msup> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 11
<p>Lift coefficient (<b>a</b>), drag coefficient (<b>b</b>), and aerodynamic efficiency (<b>c</b>) variation as a function of the jet angle; the vertical bars characterize the corresponding peak-to-peak amplitude. Constant momentum coefficient (<math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>μ</mi> </msub> <mo>=</mo> <mn>0.007</mn> </mrow> </semantics></math>) and forcing frequency (<math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mo>+</mo> </msup> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 12
<p>Lift coefficient (<b>a</b>), drag coefficient (<b>b</b>), and aerodynamic efficiency (<b>c</b>) variation as a function of forcing frequency; the vertical bars characterize the corresponding peak-to-peak amplitude. Constant momentum coefficient (<math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>μ</mi> </msub> <mo>=</mo> <mn>0.007</mn> </mrow> </semantics></math>) and jet angle (<math display="inline"><semantics> <mrow> <msub> <mi>θ</mi> <mi>jet</mi> </msub> <mo>=</mo> <mn>5</mn> <mspace width="0.222222em"/> </mrow> </semantics></math> deg.).</p>
Full article ">Figure 13
<p>Streamlines of temporal average velocity field and contours of averaged pressure in Section 54 for the baseline case (<b>a</b>) and optimum AFC implementation case (<b>b</b>).</p>
Full article ">Figure 14
<p>Streamlines of temporal average velocity field and contours of non-dimensional turbulence viscosity in Section 54 for the baseline case (<b>a</b>) and maximum lift case (<b>b</b>).</p>
Full article ">Figure 15
<p>Instantaneous velocity streamlines and non-dimensional turbulence viscosity at different phases of the synthetic jet for the maximum lift case in Section 54. <math display="inline"><semantics> <mrow> <msup> <mi>F</mi> <mo>+</mo> </msup> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>C</mi> <mi>μ</mi> </msub> <mo>=</mo> <mn>0.007</mn> <mo>,</mo> <mi>θ</mi> <mo>=</mo> <msup> <mn>5</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 16
<p>Pressure and tangential wall shear stress coefficients between the baseline case and the optimal AFC implementation on Section 54.</p>
Full article ">Figure 17
<p>Mean velocity profiles tangent to the airfoil surface at several streamwise locations, from <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>C</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> to <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>C</mi> <mo>=</mo> <mn>0.99</mn> </mrow> </semantics></math>, and as a function of the normalized wall normal distance. Comparison of the baseline with the optimum AFC case.</p>
Full article ">
18 pages, 5721 KiB  
Article
A Novel Simulation Model of Shielding Performance Based on the Anisotropic Magnetic Property of Magnetic Shields
by Yuzheng Ma, Minxia Shi, Leran Zhang, Teng Li, Xuechen Ling, Shuai Yuan, Hanxing Wang and Yi Gao
Materials 2024, 17(23), 5906; https://doi.org/10.3390/ma17235906 - 2 Dec 2024
Viewed by 422
Abstract
To achieve a near-zero magnetic field environment, the use of permalloy sheets with high-performance magnetic properties is essential. However, mainstream welding processes for magnetically shielded rooms (MSRs), such as argon arc welding and laser welding, can degrade the magnetic properties of the material. [...] Read more.
To achieve a near-zero magnetic field environment, the use of permalloy sheets with high-performance magnetic properties is essential. However, mainstream welding processes for magnetically shielded rooms (MSRs), such as argon arc welding and laser welding, can degrade the magnetic properties of the material. Additionally, neglecting the anisotropy of permalloy sheets can introduce unpredictable errors in the evaluation of MSR performance. To address this issue, this paper proposes a modified model for calculating the shielding factor (SF) of MSRs that incorporates the anisotropic magnetic characteristics of permalloy sheets. These characteristics were measured using a two-dimensional single sheet tester (2D-SST). A high-precision measurement system was developed, comprising a 2D-SST (to generate two-dimensional magnetic fields and sense the induced B and H signals) and a control system (to apply in-phase 2D excitation signals and amplify, filter, and record the B and H data). Hysteresis loops were tested at low frequencies (0.1–9 Hz) and under different magnetization states (0.1–0.6 T) in two orientations—parallel and perpendicular to the annealing magnetic field—to verify anisotropy under varying conditions. Initial permeability, near-saturation magnetization, and basic magnetization curves (BM curves) were measured across different directions to provide parameters for simulations and theoretical calculations. Based on these measurements and finite element simulations, a mathematical model was developed to adjust the empirical coefficient λ used in theoretical SF calculations. The results revealed that the ratio of empirical coefficients in different directions is inversely proportional to the ratio of magnetic permeability in the corresponding directions. A verification group was established to compare the original model and the modified model. The mean squared error (MSE) between the original model and the finite element simulation was 49.97, while the MSE between the improved model and the finite element simulation was reduced to 0.13. This indicates a substantial improvement in the computational accuracy of the modified model. Full article
Show Figures

Figure 1

Figure 1
<p>The desired field and the feedback measurement system: (<b>a</b>) desired field; (<b>b</b>) schematic of the experimental measurement system; (<b>c</b>) the physical single sheet tester (SST).</p>
Full article ">Figure 2
<p>The prototype of the 2D-SST measurement system.</p>
Full article ">Figure 3
<p>The anisotropic hysteresis loop of B = 0.1~0.7. (<b>a</b>) Argon arc welding—parallel to magnetic annealing direction, (<b>b</b>) argon arc welding—perpendicular to magnetic annealing direction, (<b>c</b>) laser welding—parallel to magnetic annealing direction, (<b>d</b>) laser welding—perpendicular to magnetic annealing direction.</p>
Full article ">Figure 4
<p>The anisotropic hysteresis loops at low-frequency band. (<b>a</b>) Argon arc welding—parallel to magnetic annealing direction, (<b>b</b>) argon arc welding—perpendicular to magnetic annealing direction, (<b>c</b>) laser welding—parallel to magnetic annealing direction, (<b>d</b>) laser welding—perpendicular to magnetic annealing direction.</p>
Full article ">Figure 5
<p>Magnetized states in various directions under excitation magnetic fields of 20 A/m and 40 A/m at 9 Hz. (<b>a</b>) Argon arc welding. (<b>b</b>) Laser welding.</p>
Full article ">Figure 6
<p>Fundamental magnetization curve. (<b>a</b>) Argon arc welding. (<b>b</b>) Laser welding.</p>
Full article ">Figure 7
<p>Anisotropic initial permeability.</p>
Full article ">Figure 8
<p>Simulation results of magnetic field distribution. (<b>a</b>–<b>c</b>) Isotropic permeability μ = 20,000 -xoy plane, yoz plane, and zox plane. (<b>d</b>–<b>f</b>) Argon arc welding—xoy plane, yoz plane, and zox plane. (<b>g</b>–<b>i</b>) Laser welding—xoy plane, yoz plane, and zox plane.</p>
Full article ">Figure 9
<p>Fluctuation range of magnetic field and origin field strength. (<b>a</b>) Isotropic permeability. (<b>b</b>) Anisotropic permeability.</p>
Full article ">Figure 10
<p>Comparison of theoretical calculated values with finite element simulation for SF. (<b>a</b>) Isotropic permeability. (<b>b</b>) Anisotropic permeability—argon arc welding. (<b>c</b>) Anisotropic permeability—lasers welding.</p>
Full article ">Figure 11
<p>Flow chart of the calculation of empirical coefficients <span class="html-italic">λ<sub>x</sub></span>, <span class="html-italic">λ<sub>y</sub></span>, and <span class="html-italic">λ<sub>z</sub></span>.</p>
Full article ">Figure 12
<p>Linear fitting results of <span class="html-italic">λ<sub>x</sub></span>, <span class="html-italic">λ<sub>y</sub></span>, and <span class="html-italic">λ<sub>z</sub></span>.</p>
Full article ">
21 pages, 28873 KiB  
Article
High-Resolution Nearshore Sea Surface Temperature from Calibrated Landsat Brightness Data
by William H. Speiser and John L. Largier
Remote Sens. 2024, 16(23), 4477; https://doi.org/10.3390/rs16234477 - 28 Nov 2024
Viewed by 426
Abstract
Understanding and monitoring nearshore environments is essential, given that these fine-scaled ecosystems are integral to human well-being. While satellites offer an opportunity to gain synchronous and spatially extensive data of coastal areas, off-the-shelf calibrated satellite sea surface temperature (SST) measurements have only been [...] Read more.
Understanding and monitoring nearshore environments is essential, given that these fine-scaled ecosystems are integral to human well-being. While satellites offer an opportunity to gain synchronous and spatially extensive data of coastal areas, off-the-shelf calibrated satellite sea surface temperature (SST) measurements have only been available at coarse resolutions of 1 km or larger. In this study, we develop a novel methodology to create a simple linear equation to calibrate fine-scale Landsat thermal infrared radiation brightness temperatures (calibrated for land sensing) to derive SST at a resolution of 100 m. The constants of this equation are derived from correlations of coincident MODIS SST and Landsat data, which we filter to find optimal pairs. Validation against in situ sensor data at varying distances from the shore in Northern California shows that our SST estimates are more accurate than prior off-the-shelf Landsat data calibrated for land surfaces. These fine-scale SST estimates also demonstrate superior accuracy compared with coincident MODIS SST estimates. The root mean square error for our minimally filtered dataset (n = 557 images) ranges from 0.76 to 1.20 °C with correlation coefficients from r = 0.73 to 0.92, and for our optimal dataset (n = 229 images), the error is from 0.62 to 0.98 °C with correlations from r = 0.83 to 0.92. Potential error sources related to stratification and seasonality are examined and we conclude that Landsat data represent skin temperatures with an error between 0.62 and 0.73 °C. We discuss the utility of our methodology for enhancing coastal monitoring efforts and capturing previously unseen spatial complexity. Testing the calibration methodology on Landsat images before and after the temporal bounds of accurate MODIS SST measurements shows successful calibration with lower errors than the off-the-shelf, land-calibrated Landsat product, extending the applicability of our approach. This new approach for obtaining high-resolution SST data in nearshore waters may be applied to other upwelling regions globally, contributing to improved coastal monitoring, management, and research. Full article
Show Figures

Figure 1

Figure 1
<p>Map of study region in WRS path 45 row 33. (<b>Left</b>) Map of in situ validation sites, from north to south: NOAA N46014 (buoy), BOON Intake (seawater intake), BML Mooring (buoy), NOAA N46013 (buoy), and BOON Tomales Bay (buoy). (<b>Right</b>) Map of geographical points of interest, from north to south: Mendocino, Manchester Beach, Pt. Arena, Gualala River Estuary, Russian River Estuary, Salmon Creek Beach, Bodega Bay, Tomales Bay, and Pt. Reyes National Seashore. Bathymetric basemap data were accessed from the General Bathymetric Chart of the Ocean (<a href="https://www.gebco.net/" target="_blank">https://www.gebco.net/</a>), with color scaled to depth in meters as shown in the legend.</p>
Full article ">Figure 2
<p>Box and whisker plot of the number of pixels per image by month in the “all” dataset. Box margins are the 25th percentile of data (lower) and 75th percentile of data (upper). The line in the box is the monthly median. Whiskers show minimum and maximum values.</p>
Full article ">Figure 3
<p>Example of calibration of Landsat B<sub>t</sub> to SST using coincidental MODIS Terra SST data from images captured on 7 January 2015 (MODIS at 18:20 LST and Landsat at 18:51 LST). (<b>a</b>) Initial correlation between MODIS SST pixels and respective pixel area median Landsat B<sub>t</sub> pixel values from 7 January 2015. (<b>b</b>) Final correlation after iterative outlier removal as outlined in <a href="#sec2dot3-remotesensing-16-04477" class="html-sec">Section 2.3</a>. (<b>c</b>) MODIS Terra SST image from 7 January 2015. (<b>d</b>) Estimated Landsat SST obtained by calibrating Landsat B<sub>t</sub> values with the best-fit linear equation from panel (<b>b</b>).</p>
Full article ">Figure 4
<p>Monthly box and whisker plots from calibration of all data. Box margins are lower 25th percentile of data (lower) and 75th percentile of data (upper). Orange line in box is monthly median. Whiskers show minimum and maximum values. Red dotted line in c. and d. is overall average value. (<b>a</b>) Pearson’s r correlation values from first iteration of per-pixel comparisons between Landsat B<sub>t</sub> and M<sub>SST</sub>. (<b>b</b>) Pearson’s r correlation values from final iteration of per-pixel comparisons after iterative outlier removal between Landsat B<sub>t</sub> and M<sub>SST</sub>. (<b>c</b>) Intercept values from OLS best-fit linear equations from final iteration comparisons between Landsat B<sub>t</sub> and M<sub>SST</sub>. (<b>d</b>) Intercept values from OLS best-fit linear equations from final iteration comparisons between Landsat B<sub>t</sub> and M<sub>SST</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Histogram of intercept (b) values from per-image calibrations of R<sub>0</sub> data. (<b>b</b>) Histogram of slope (m) values from per-image calibrations of R<sub>0</sub> data. Red lines indicate distribution medians (b = −105.88; m = 0.00297).</p>
Full article ">Figure 6
<p>SST estimates from per-image-calibrated Landsat data plotted against water temperature from in situ sensors: (<b>a</b>) N46014, (<b>b</b>) N46013, (<b>c</b>) BML Mooring, (<b>d</b>) TB Mooring, and (<b>e</b>) BML Intake. Yellow data and trendline are for the All data dataset and red data and trendline are for R<sub>0</sub> data.</p>
Full article ">Figure 7
<p>SST estimates from generalized-equation-calibrated Landsat data plotted against water temperature measured by in situ sensors: (<b>a</b>) N46014, (<b>b</b>) N46013, (<b>c</b>) BML Mooring, (<b>d</b>) TB Mooring, and (<b>e</b>) BML Intake. Yellow data and trendline are for the All data dataset, red data and trendline are for R<sub>0</sub> data, green data and trend line are for USGS B<sub>t</sub> data and purple is MODIS SST data.</p>
Full article ">Figure 8
<p>Monthly box and whisker plots of difference between estimated SST and in situ measurements across “all data”. Box margins are lower 25th percentile of difference (lower bound) and 75th percentile of difference (upper bound). Orange line in box is monthly median. Whiskers show minimum and maximum difference values. (<b>a</b>) N46014, (<b>b</b>) N46013, (<b>c</b>) BML Mooring, (<b>d</b>) TB Mooring, (<b>e</b>) BML Intake, and (<b>f</b>) all buoy locations combined.</p>
Full article ">Figure 9
<p>In situ sensor data on temperature at (<b>a</b>) N46014, (<b>b</b>) N46013, and BML Intake (<b>c</b>) plotted against SST data from general-equation calibration of Landsat B<sub>t</sub> data before and after MODIS capture window. Green data points are data captured after 2023, and red data points are data captured prior to 2000. Blue trend line is OLS best-fit line for combination of those two data ranges.</p>
Full article ">Figure 10
<p>Mapped temperature from 25 April 2019 vs. true-color values at zoomed-in sites. Color scale adjusted to highlight all temperature features. (<b>a</b>) Entire calibrated SST image. (<b>b</b>) Zoomed-in calibrated SST (left) and true-color imagery (right) directly west of the Russian River estuary. (<b>c</b>) Zoomed-in calibrated SST (left) and true-color imagery (right) north of the Pt. Reyes National Seashore and south of Bodega Bay. (<b>d</b>) Zoomed-in calibrated SST (left) and true-color imagery (right) of the north headland of Point Reyes national seashore.</p>
Full article ">Figure 11
<p>Cross-shore temperature profiles from a transect at Manchester Beach, CA, extending 6.5 km offshore from per-image calibrated images. (<b>Top</b>) Plots of temperature (y-axis) vs. distance (x-axis). (<b>Bottom</b>) Mapped temperature values from dates respective to plots in the same column.</p>
Full article ">Figure 12
<p>Average SST at each pixel position in calibrated Landsat data from 1983 to 2024 in degrees Celsius calibrated using generalized equation. Isothermal contours are shown for 11.5°, 12.0°, 12.5°, and 13.0 °C.</p>
Full article ">
20 pages, 11913 KiB  
Article
Long-Term Spatiotemporal Analysis of Precipitation Trends with Implications of ENSO-Driven Variability in the Department of Magdalena, Colombia
by Geraldine M. Pomares-Meza, Yiniva Camargo Caicedo and Andrés M. Vélez-Pereira
Water 2024, 16(23), 3372; https://doi.org/10.3390/w16233372 - 23 Nov 2024
Viewed by 694
Abstract
The Magdalena department, influenced by southern trade winds and ocean currents from the Atlantic and Pacific, is a climatically vulnerable region. This study assesses the Magdalena Department’s precipitation trends and stationary patterns by analyzing multi-year monthly records from 55 monitoring stations from 1990 [...] Read more.
The Magdalena department, influenced by southern trade winds and ocean currents from the Atlantic and Pacific, is a climatically vulnerable region. This study assesses the Magdalena Department’s precipitation trends and stationary patterns by analyzing multi-year monthly records from 55 monitoring stations from 1990 to 2022. To achieve this, the following methods were used: (i) homogeneous regions were established by an unsupervised clustering approach, (ii) temporal trends were quantified using non-parametric tests, (iii) stationarity was identified through Morlet wavelet decomposition, and (iv) Sea Surface Temperature (SST) in four Niño regions was correlated with stationarity cycles. Silhouette’s results yielded five homogeneous regions, consistent with the National Meteorological Institute (IDEAM) proposal. The Department displayed decreasing annual trends (−32–−100 mm/decade) but exhibited increasing monthly trends (>20 mm/decade) during the wettest season. The wavelet decomposition analysis revealed quasi-bimodal stationarity, with significant semiannual cycles (~4.1 to 5.6 months) observed only in the eastern region. Other regions showed mixed behavior: non-stationary in the year’s first half and stationary in the latter half. Correlation analysis showed a significant relationship between SST in the El Niño 3 region (which accounted for 50.5% of the coefficients), indicating that strong phases of El Niño anticipated precipitation responses for up to six months. This confirms distinct rainfall patterns and precipitation trends influenced by the El Niño–Southern Oscillation (ENSO), highlighting the need for further hydrometeorological research in the area. Full article
Show Figures

Figure 1

Figure 1
<p>General location, physiographic units, and hydrometeorological station network of the Department of Magdalena (map on the left), detailing the spatial distribution of average annual precipitation in the Department of Magdalena (1990–2022) (map on the right).</p>
Full article ">Figure 2
<p>The conceptual framework of the proposed methodology.</p>
Full article ">Figure 3
<p>Comparison between clustering scenarios. The size of the points is proportional to the average Silhouette coefficient (average cluster performance), and the transparency level indicates the individual Silhouette coefficient (station affinity to the cluster’s centroid). Sites indicated by empty symbols represent the centroid for each cluster. (<b>A</b>) Euclidean + non-standardization scenario. (<b>B</b>) Euclidean + z-score scenario (selected scenario). (<b>C</b>) DTW + non-standardization scenario. (<b>D</b>) DTW + z-score scenario.</p>
Full article ">Figure 4
<p>Spatial distribution of total annual rainfall in the final configuration of homogeneous regions in the Magdalena department.</p>
Full article ">Figure 5
<p>Results by type of trend (direction of the triangle), statistical significance (size of the triangle), and value of the magnitude of change in mm decade<sup>−1</sup> (color scale) of total annual rainfall (map on the left) and total monthly rainfall (small maps on the right) in the Magdalena department.</p>
Full article ">Figure 6
<p>Results by type of trend (direction of the triangle), statistical significance (size of the triangle), and value of the magnitude of change in day/decade (color scale) of total annual rainy days (map on the left) and total monthly rainy days (small maps on the right) in the Magdalena department.</p>
Full article ">Figure 7
<p>Distribution of wavelet power across homogeneous rainfall regions in the Magdalena department. For all scalograms, the x-axis represents the time component, and the y-axis represents the scale component, whose limits were adjusted from a 3- to a 120-month (10-year) scale. The color scale indicates wavelet power variation and represents each time-scale component’s contribution to the rainfall series’ variance. Black crosses delimit the regions of significant stationarity, calculated using the Torrence and Compo [<a href="#B72-water-16-03372" class="html-bibr">72</a>] proposed test at a 5% significance level.</p>
Full article ">Figure 8
<p>Wavelet coherence (<b>left</b>) and phase difference (<b>right</b>) between multi-year precipitation and average SST in the El Niño 3 region. For all scalograms, the x-axis represents the time component (1990–2022), and the y-axis represents the scale component, whose limits were adjusted from a 3- to 120-month (i.e., 10-year) scale. The color scale indicates wavelet coherence and phase difference variation and represents both parameters’ correlation and synchronization (respectively) at the specific time-scale component. Significant correlation at a 5% significance level is represented by dashed black lines, indicating the periods most likely influenced by SST seasonality in their corresponding cycle length.</p>
Full article ">Figure 9
<p>Comparison between (<b>A</b>) Thornthwaite moisture index classification proposed by IDEAM (2017) and (<b>B</b>) clustering-based precipitation regionalization in the department of Magdalena.</p>
Full article ">
21 pages, 3854 KiB  
Article
Optimization of a Gorlov Helical Turbine for Hydrokinetic Application Using the Response Surface Methodology and Experimental Tests
by Juan Camilo Pineda, Ainhoa Rubio-Clemente and Edwin Chica
Energies 2024, 17(22), 5747; https://doi.org/10.3390/en17225747 - 17 Nov 2024
Viewed by 515
Abstract
The work presents an analysis of the Gorlov helical turbine (GHT) design using both computational fluid dynamics (CFD) simulations and response surface methodology (RSM). The RSM method was applied to investigate the impact of three geometric factors on the turbine’s power coefficient (C [...] Read more.
The work presents an analysis of the Gorlov helical turbine (GHT) design using both computational fluid dynamics (CFD) simulations and response surface methodology (RSM). The RSM method was applied to investigate the impact of three geometric factors on the turbine’s power coefficient (CP): the number of blades (N), helix angle (γ), and aspect ratio (AR). Central composite design (CCD) was used for the design of experiments (DOE). For the CFD simulations, a three-dimensional computational domain was established in the Ansys Fluent software, version 2021R1 utilizing the k-ω SST turbulence model and the sliding mesh method to perform unsteady flow simulations. The objective function was to achieve the maximum CP, which was obtained using a high-correlation quadratic mathematical model. Under the optimum conditions, where N, γ, and AR were 5, 78°, and 0.6, respectively, a CP value of 0.3072 was achieved. The optimal turbine geometry was validated through experimental testing, and the CP curve versus tip speed ratio (TSR) was determined and compared with the numerical results, which showed a strong correlation between the two sets of data. Full article
(This article belongs to the Section B: Energy and Environment)
Show Figures

Figure 1

Figure 1
<p>Geometric factors involved in the Gorlov helical turbine (GHT) design.</p>
Full article ">Figure 2
<p>Dimensions of the computational domain and the setup of boundary conditions for the Ansys Fluent simulation.</p>
Full article ">Figure 3
<p>Mesh of computational domain dimensions.</p>
Full article ">Figure 4
<p>Experimental setup of the recirculating water channel. (1) Motor of 14.9 kW, (2) impeller, (3) water inlet value, (4) channel, (5) gate, (6) model vertical-axis hydrokinetic turbine, (7) connection axis to the sensor, (8) weir assembly, and (9) feed tank.</p>
Full article ">Figure 5
<p>Response surface plots for the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) by using the regression model. (<b>a</b>) Effects of N and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>; (<b>b</b>) effects of N and AR; (<b>c</b>) effects of <math display="inline"><semantics> <mi>γ</mi> </semantics></math> and AR. The other factors were set at the optimal values.</p>
Full article ">Figure 6
<p>(<b>a</b>) Frequency distribution and (<b>b</b>) normal probability plots for the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>).</p>
Full article ">Figure 7
<p>Numerical and experimental comparison of the power coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) vs. tip speed ratio (<math display="inline"><semantics> <mi>λ</mi> </semantics></math>) curves.</p>
Full article ">
15 pages, 4870 KiB  
Article
Research on Effect of Ship Speed on Unsteady Hydrodynamic Performance of Bow Thrusters in Berthing and Departure Directions
by He Cai, Xiaoqian Ma, Tan Wen, Yu Sun, Zhiyuan Yang, Yilong Tan and Jianyu Zhuo
J. Mar. Sci. Eng. 2024, 12(11), 2054; https://doi.org/10.3390/jmse12112054 - 13 Nov 2024
Viewed by 506
Abstract
With the continuous development of the shipping market, bow thrusters have become more important for ship maneuvering. Therefore, the performance of bow thrusters is studied in this paper. In order to obtain an unsteady performance of the bow thruster under different ship speed [...] Read more.
With the continuous development of the shipping market, bow thrusters have become more important for ship maneuvering. Therefore, the performance of bow thrusters is studied in this paper. In order to obtain an unsteady performance of the bow thruster under different ship speed conditions, the SST k-ω turbulence model is adopted to predict the hydrodynamics of the bow thruster. With the ship’s speed increasing gradually, the variation characteristics of hydrodynamic coefficients and the flow field distribution at key positions are analyzed. The results show that with an increase in ship speed to three knots, the thrust coefficient and torque coefficient of the bow thruster decrease by 2.69~4.07% and 2.34~3.08%. In addition, the blade vibration amplitude intensifies. In the departure direction, the propeller load is more susceptible to being influenced and decreases by an additional 2.34~4.16% compared with that in the berthing direction. Meanwhile, it is found that the velocity distribution is asymmetrical. The inlet velocity at the bow side is faster, which results in the maximum peak pressure being about three times the minimum peak pressure. In addition, the pressure’s nonuniformity in the tunnel increases gradually with the increase in ship speed. Compared with the pressure distribution in the berthing direction, the pressure distribution before and after the propeller is more uniform, which is consistent with the results of hydrodynamic change and velocity distribution. The research in this paper has a certain reference significance for understanding the hydrodynamic performance of bow thrust operation. Full article
Show Figures

Figure 1

Figure 1
<p>Geometric model of bow thruster.</p>
Full article ">Figure 2
<p>Hydrodynamic model of propeller and its position in tunnel.</p>
Full article ">Figure 3
<p>Hydrodynamic model of ship under waterline.</p>
Full article ">Figure 4
<p>Computational domain and boundary conditions.</p>
Full article ">Figure 5
<p>Grid scheme and <span class="html-italic">y</span><sup>+</sup> value of computational model.</p>
Full article ">Figure 6
<p>Ducted propeller with its grid scheme.</p>
Full article ">Figure 7
<p>Open water performance curves of ducted propeller.</p>
Full article ">Figure 8
<p>Pressure distribution on the blades (<b>left</b>: pressure side, <b>right</b>: suction side).</p>
Full article ">Figure 9
<p>Monitoring point positions with their peak pressures.</p>
Full article ">Figure 10
<p>Pressure pulsations at different positions on the tunnel surface.</p>
Full article ">Figure 11
<p>Thrust coefficient curves in berthing direction.</p>
Full article ">Figure 12
<p>Torque coefficient curves in berthing direction.</p>
Full article ">Figure 13
<p>Velocity distributions at propeller shaft depth in berthing direction.</p>
Full article ">Figure 14
<p>Pressure distributions at propeller shaft depth in berthing direction.</p>
Full article ">Figure 15
<p>Thrust coefficient curves in departure direction.</p>
Full article ">Figure 16
<p>Torque coefficient curves in departure direction.</p>
Full article ">Figure 17
<p>Velocity distributions at propeller shaft depth in departure direction.</p>
Full article ">Figure 18
<p>Pressure distributions at propeller shaft depth in departure direction.</p>
Full article ">
22 pages, 12960 KiB  
Article
Aerodynamic Performance and Numerical Validation Study of a Scaled-Down and Full-Scale Wind Turbine Models
by Zahid Mehmood, Zhenyu Wang, Xin Zhang and Guiying Shen
Energies 2024, 17(21), 5449; https://doi.org/10.3390/en17215449 - 31 Oct 2024
Viewed by 1670
Abstract
Understanding the aerodynamic performance of scaled-down models is vital for providing crucial insights into wind energy optimization. In this study, the aerodynamic performance of a scaled-down model (12%) was investigated. This validates the findings of the unsteady aerodynamic experiment (UAE) test sequence H. [...] Read more.
Understanding the aerodynamic performance of scaled-down models is vital for providing crucial insights into wind energy optimization. In this study, the aerodynamic performance of a scaled-down model (12%) was investigated. This validates the findings of the unsteady aerodynamic experiment (UAE) test sequence H. UAE tests provide information on the configuration and conditions of wind tunnel testing to measure the pressure coefficient distribution on the blade surface and the aerodynamic performance of the wind turbine. The computational simulations used shear stress transport and kinetic energy (SST K-Omega) and transitional shear stress transport (SST) turbulence models, with wind speeds ranging from 5 m/s to 25 m/s for the National Renewable Energy Laboratory (NREL) Phase VI and 4 m/s to 14 m/s for the 12% scaled-down model. The aerodynamic performance of both cases was assessed at representative wind speeds of 7 m/s for low, 10 m/s for medium, and 20 m/s for high flow speeds for NREL Phase VI and 7 m/s for low, 9 m/s medium, and 12 m/s for the scaled-down model. The results of the SST K-Omega and transitional SST models were aligned with experimental test measurement data at low wind speeds. However, the SST K-Omega torque values exhibited a slight deviation. The transitional SST and SST K-Omega models yielded aerodynamic properties that were comparable to those of the 12% scaled-down model. The torque values obtained from the simulation for the full-scale NREL Phase VI and the scaled-down model were 1686.5 Nm and 0.8349 Nm, respectively. Both turbulence models reliably predicted torque and pressure coefficient values that were consistent with the experimental data, considering specific flow regimes. The pressure coefficient was maximum at the leading edge of the wind turbine blade on the windward side and minimum on the leeward side. For the 12% scaled-down model, the flow simulation results bordering the low-pressure region of the blade varied slightly. Full article
Show Figures

Figure 1

Figure 1
<p>Details of the domain used for the CFD simulation.</p>
Full article ">Figure 2
<p>Meshing for simulations.</p>
Full article ">Figure 3
<p>Comparison of torque (<b>a</b>) and thrust force (<b>b</b>) using CFD and experimental data for NREL Phase VI.</p>
Full article ">Figure 4
<p>Comparison of torque (<b>a</b>) and thrust force (<b>b</b>) using CFD and experimental data for the scaled-down model based on NREL Phase VI.</p>
Full article ">Figure 5
<p>Comparison of the coefficient of pressure of NREL Phase VI obtained from CFD simulations and experimental data at a free-stream velocity of 7 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 6
<p>Comparison of the coefficient of pressure of NREL Phase VI obtained from CFD simulations and experimental data at a free-stream velocity of 10 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 7
<p>Comparison of the coefficient of pressure of NREL Phase VI obtained from CFD simulations and experimental data at a free-stream velocity of 20 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 8
<p>Comparison of pressure coefficients obtained from transitional SST and SST K-Omega turbulence models at a wind velocity of 7 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 9
<p>Pressure variation contours at different radial locations of the blade by SST K-Omega (graph with title SSTKP) and transitional SST (graph with title TSSTP) models at a wind velocity of =7 m/s.</p>
Full article ">Figure 10
<p>Comparison of pressure coefficients obtained from transitional SST and SST K-Omega turbulence models at a wind speed of 9 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 11
<p>Pressure variation contours at different radial locations of the blade using SST K-Omega (graph with title SSTKP) and transitional SST (graph with title TSSTP) models at a wind speed of 9 m/s.</p>
Full article ">Figure 12
<p>Comparison of pressure coefficients obtained from SST K-Omega and transitional SST models at a wind speed of 12 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 12 Cont.
<p>Comparison of pressure coefficients obtained from SST K-Omega and transitional SST models at a wind speed of 12 m/s and at different blade sections (<b>a</b>–<b>e</b>).</p>
Full article ">Figure 13
<p>Pressure variation contours at different radial locations of the blade using SST K-Omega (graph with title SSTKP) and transitional SST (graph with title TSSTP) models at a wind speed of 12 m/s.</p>
Full article ">Figure 13 Cont.
<p>Pressure variation contours at different radial locations of the blade using SST K-Omega (graph with title SSTKP) and transitional SST (graph with title TSSTP) models at a wind speed of 12 m/s.</p>
Full article ">
14 pages, 11753 KiB  
Article
Wear Behaviour of Graphene-Reinforced Ti-Cu Waste-Metal Friction Composites Fabricated with Spark Plasma Sintering
by Mária Podobová, Viktor Puchý, Richard Sedlák, Dávid Medveď, Róbert Džunda and František Kromka
Crystals 2024, 14(11), 948; https://doi.org/10.3390/cryst14110948 - 31 Oct 2024
Viewed by 574
Abstract
In this study, we fabricated Ti-Cu-based friction composites containing waste-metal (Ti, CuZn, stainless steel (SSt), MgAl), Al2O3 due to improving properties and its good compatibility with copper and graphene nanoplatelets as reinforcement and lubricant component, using planetary ball mill and [...] Read more.
In this study, we fabricated Ti-Cu-based friction composites containing waste-metal (Ti, CuZn, stainless steel (SSt), MgAl), Al2O3 due to improving properties and its good compatibility with copper and graphene nanoplatelets as reinforcement and lubricant component, using planetary ball mill and technique based on Spark Plasma Sintering (SPS). Understanding the wear behaviour of such engineered friction composites is essential to improve their material design and safety, as these materials could have the potential for use in public and industrial transportation, such as high-speed rail trains and aircraft or cars. This is why our study is focused on wear behaviour during friction between function parts of devices. We investigated the composite materials designed by us in order to clarify their microstructural state and mechanical properties. Using different loading conditions, we determined the Coefficient of Friction (COF) using a ball-on-disc tribological test. We analysed the state of the samples after the mentioned test using a Scanning Electron Microscope (SEM), then Energy-Dispersive X-ray Spectroscopy (EDS), and confocal microscopy. Also, a comparative analysis of friction properties with previously studied materials was performed. The results showed that friction composites with different compositions, despite the same conditions of their compaction during sintering, can be defined by different wear characteristics. Our study can potentially have a significant contribution to the understanding of wear mechanisms of Ti-Cu-based composites with incorporated metal-waste and to improving their material design and performance. Also, it can give us information about the possibilities of reusing metal-waste from different machining operations. Full article
(This article belongs to the Special Issue Processing, Structure and Properties of Metal Matrix Composites)
Show Figures

Figure 1

Figure 1
<p>Morphologies of the four input waste materials: (<b>a</b>) Ti, (<b>b</b>) MgAl, (<b>c</b>) CuZn, (<b>d</b>) Cu, and (<b>e</b>) SSt stainless steel.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) TC1–TC3 (from left to right) composite powder mixtures after planetary ball milling.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Sintering curves, shrinkage over time, and applied pressure force.</p>
Full article ">Figure 4
<p>Coefficient of friction vs. time for different values of load (5 N, 10 N).</p>
Full article ">Figure 5
<p>Average coefficient of friction for applied load 5N and 10N with deviation values.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) SEM-EDX mapping TC1–TC3 area (<b>from left to right</b>).</p>
Full article ">Figure 7
<p>SEM from wear track TC1–TC3 (<b>from left to right</b>), (<b>a</b>–<b>c</b>) load 5 N, (<b>d</b>–<b>f</b>) load 10 N.</p>
Full article ">Figure 8
<p>Visualisation of the confocal images of wear tracks (<b>a</b>–<b>c</b>) load 5 N, (<b>d</b>–<b>f</b>) load 10 N.</p>
Full article ">Figure 9
<p>XRD records of samples TC1–TC3 (<b>from left to right</b>).</p>
Full article ">
27 pages, 12462 KiB  
Article
Long-Term Teleconnections Between Global Circulation Patterns and Interannual Variability of Surface Air Temperature over Kingdom of Saudi Arabia
by Abdullkarim K. Almaashi, Hosny M. Hasanean and Abdulhaleem H. Labban
Atmosphere 2024, 15(11), 1310; https://doi.org/10.3390/atmos15111310 - 30 Oct 2024
Viewed by 469
Abstract
Surface air temperature (SAT) variability is investigated for advancing our understanding of the climate patterns over the Kingdom of Saudi Arabia (KSA). SAT variability reveals significant warming trends, particularly from 1994 onward, as demonstrated by nonlinear and linear trend analysis. This warming is [...] Read more.
Surface air temperature (SAT) variability is investigated for advancing our understanding of the climate patterns over the Kingdom of Saudi Arabia (KSA). SAT variability reveals significant warming trends, particularly from 1994 onward, as demonstrated by nonlinear and linear trend analysis. This warming is linked to global climate patterns, which serve as significant indicators for studying the effects of climate change on surface air temperature patterns across the KSA. The empirical orthogonal function (EOF) method is employed for analyzing SAT due to its effectiveness in extracting dominant patterns of variability during the winter (DJF) and summer (JJA) seasons. The first mode (EOF1) for both seasons shows positive variability across the KSA, explaining more than 45% of the variance. The second mode (EOF2) indicates negative variability in central and northern regions. The third mode (EOF3) describes positive variability but with lower variance over time. PC1 is used to describe the physical mechanism of SAT variability and correlations with global sea surface temperature (SST). The physical mechanism shows that the variability in Mediterranean troughs during the winter season and high pressure over the Indian Ocean and central Asia controls SAT variability over the KSA. The correlation coefficients (CCs) were calculated during the winter and summer season between the SAT of the KSA and six teleconnection indices, El Niño Southern Oscillation (ENSO), Indian Ocean Dipole (IOD), Atlantic Meridional Mode (AMM), Pacific Warm Pool (PWP), North Atlantic Oscillation (NAO), and Tropical North Atlantic (TNA) SST for the period from 1994 to 2022. ENSO shifts from positive to negative correlations with SAT from winter to summer. IOD shows a diminished correlation with SAT due to the absence of upper air dynamics. PWP consistently enhances surface warming in both seasons through upper air convergence during both seasons. AMM and NAO have a non-significant impact on SAT; however, TNA contributes warming over central and northern parts during winter and summer seasons. The seasonal SAT variations emphasize the significant role of ENSO, PWP, and TNA across the seasons. The findings of this study can be helpful for seasonal predictability in the KSA. Full article
(This article belongs to the Section Climatology)
Show Figures

Figure 1

Figure 1
<p>The elevation in meters, along with solid circles representing the observation station names in Saudi Arabia. (Data source: USGS satellite topographic dataset).</p>
Full article ">Figure 2
<p>Analysis of the changes in ERA5 SAT over the KSA for (<b>a</b>) the summer season (June to August), (<b>b</b>) the winter season (December–January and February). The red lines represent the mean SAT scores for the times 1952–1993 and 1994–2022. Global analysis of the changes in ERA5 SAT of (<b>c</b>) the summer season (December–January and February) and (<b>d</b>) winter season (June to August).</p>
Full article ">Figure 3
<p>(<b>a</b>) Summer season trend magnitude (Sen’s slope °C/year) during 1952–1993, (<b>b</b>) same as (<b>a</b>) but for the period 1994–2022. (<b>c</b>) Winter season trend magnitude (Sen’s slope °C/year) during 1952–1993, (<b>d</b>) same as (<b>c</b>) but for the period 1994–2022. Dotted areas show the significance above 99% confidence level.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>): The Empirical Orthogonal Function (EOF) analysis of SAT temperature for Jun–Aug season during 1952–1993. (<b>d</b>–<b>f</b>) show the corresponding PCs of EOF analyses.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) The Empirical Orthogonal Function (EOF) analysis of SAT temperature for Jun–Aug season during 1994–2022. (<b>d</b>–<b>f</b>) show the corresponding PCs of EOFs.</p>
Full article ">Figure 6
<p>(<b>a</b>) The leading EOF mode represents the primary pattern, capturing 64.70% of the total variance. (<b>b</b>) The second EOF mode follows, explaining 14.51% of the variance. (<b>c</b>) The third EOF mode contributes to 5.75% of the total variance. (<b>d</b>–<b>f</b>) The associated principal component time series (PC1, PC2, and PC3) correspond to these leading EOF modes.</p>
Full article ">Figure 7
<p>(<b>a</b>) The leading EOF mode represents the primary pattern, capturing 54.26% of the total variance. (<b>b</b>) The second EOF mode follows, explaining 19.07% of the variance. (<b>c</b>) The third EOF mode contributes to 7.22% of the total variance. (<b>d</b>–<b>f</b>) The associated principal component time series (PC1, PC2, and PC3) correspond to these leading EOF modes.</p>
Full article ">Figure 8
<p>Correlation between PC1 of SAT and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during winter (<b>a</b>) for 1952–1993, (<b>b</b>) for 1994–2022. (<b>c,d</b>) Same as (<b>a</b>,<b>b</b>) but for summer season. The dotted areas show the regions with significance level above 95% by using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 9
<p>(<b>a</b>) Correlation between SAT and ENSO, computed from standardized anomalies of winter seasonal time series spanning 1952–2022. Dotted regions indicate significance levels exceeding 95% confidence, determined using Student’s <span class="html-italic">t</span>-test. (<b>b</b>–<b>f</b>) Same as (<b>a</b>) but for the IOD, PWP, AMM, NAO, and TNA.</p>
Full article ">Figure 10
<p>(<b>a</b>) Correlation between SAT and ENSO, computed from standardized anomalies of summer seasonal time series spanning 1952–2022. Dotted regions indicate significance levels exceeding 95% confidence, determined using Student’s <span class="html-italic">t</span>-test. (<b>b</b>–<b>f</b>) same as (<b>a</b>) but for the IOD, PWP, AMM, NAO, and TNA.</p>
Full article ">Figure 11
<p>Correlation between climate indices and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during winter from 1994 to 2022. (<b>a</b>) Nino3.4, (<b>b</b>) IOD, (<b>c</b>) PWP, (<b>d</b>) AMM, (<b>e</b>) NAO, and (<b>f</b>) TNA. The dotted areas show the regions with significance level above 95% by using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 12
<p>Correlation between climate indices and MSLP (shaded) and 200 hPa divergent wind fields (vectors) during summer season from 1994 to 2022. (<b>a</b>) Nino3.4, (<b>b</b>) IOD, (<b>c</b>) PWP, (<b>d</b>) AMM, (<b>e</b>) NAO, and (<b>f</b>) TNA. The dotted areas show the regions with significance level above 95% using Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 13
<p>The correlation coefficient map between PC1 and global SST for DJF season during (<b>a</b>) 1952–1993, (<b>b</b>) during the period 1994–2022. Stippling denotes regions where the relationship is statistically significant at 95% confidence level based on Student’s <span class="html-italic">t</span>-test.</p>
Full article ">Figure 14
<p>The correlation coefficient maps between PC1 and global SST for JJA season during (<b>a</b>) 1952–1993, (<b>b</b>) during the period 1994–2022. Stippling denotes regions where the relationship is statistically significant at 95% confidence level based on Student’s <span class="html-italic">t</span>-test.</p>
Full article ">
16 pages, 6843 KiB  
Article
Seasonal–Diurnal Distribution of Lightning over Bulgaria and the Black Sea and Its Relationship with Sea Surface Temperature
by Savka Petrova, Rumjana Mitzeva, Vassiliki Kotroni and Elisaveta Peneva
Atmosphere 2024, 15(10), 1233; https://doi.org/10.3390/atmos15101233 - 15 Oct 2024
Viewed by 461
Abstract
A seasonal–diurnal analysis of land-sea contrast in lightning activity over Bulgaria and the Black Sea over 10 years is presented here. The maximum number of flashes over both surface types is registered during the summer (with a peak over Bulgaria in June and [...] Read more.
A seasonal–diurnal analysis of land-sea contrast in lightning activity over Bulgaria and the Black Sea over 10 years is presented here. The maximum number of flashes over both surface types is registered during the summer (with a peak over Bulgaria in June and over the Black Sea in July) and a minimum number in winter (December/February, respectively). During spring, the maximum flash density is observed over Bulgaria (in May), while in autumn, it is over the Black Sea (in September). The results show that only in autumn lightning activity dominates over the Black Sea compared to over land (Bulgaria), while in winter, spring, and summer is vice versa. For this reason, an additional investigation was conducted to determine whether there is a relationship between lightning activity and the sea surface temperature (SST) of the Black Sea in autumn. The analysis reveals that the influence of SST on the formation of thunderstorms over the Black Sea varies depending on the diurnal time interval, with the effect being more significant at night. At nighttime intervals, there is a clear trend of increasing mean flash frequency per case with rising SST (linear correlation coefficients range from R = 0.92 to 0.98), while during the daytime, this trend is not as evident. This indicates that, during the day, other favorable atmospheric processes have a greater influence on the formation of thunderstorms than sea-surface temperature, while in the autumn night hours, the higher SST values probably play a more significant role in thunderstorms formation, in combination with the corresponding orographic conditions. Full article
(This article belongs to the Special Issue Atmospheric Electricity (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Elevation [m] in the studied area. Positive numbers represent the elevation above the sea level, negative numbers—below the sea level. The data are taken from the GEBCO 1-min global grid (<a href="http://www.gebco.net" target="_blank">www.gebco.net</a>, accessed on 1 May 2022).</p>
Full article ">Figure 2
<p>Spatial distribution of the number of recorded flashes within 0.25° × 0.25° grid boxes over Bulgaria and the Black Sea during winter (<b>a</b>), spring (<b>b</b>), summer (<b>c</b>), and autumn (<b>d</b>) from March 2005 to February 2015. The color scale represents the number of flashes; note that the scale is different in each season.</p>
Full article ">Figure 3
<p>Flash density (flashes/km<sup>2</sup>) over the Black Sea (blue columns) and Bulgaria (white columns) during winter (DJF), spring (MAM), summer (JJA), and autumn (SON) from March 2005 to February 2015.</p>
Full article ">Figure 4
<p>Flash density (flashes/km<sup>2</sup>) for months in winter (<b>a</b>), spring (<b>b</b>), summer (<b>c</b>), and autumn (<b>d</b>) over the Black Sea (blue columns) and Bulgaria (white columns), March 2005–February 2015.</p>
Full article ">Figure 5
<p>Diurnal variation of flash density (flashes/km<sup>2</sup>) at 3-h time intervals during winter (<b>a</b>), spring (<b>b</b>), summer (<b>c</b>), and autumn (<b>d</b>) over the Black Sea (blue columns) and Bulgaria (white columns) from March 2005 to February 2015. Local time is UTC + 2 h or UTC + 3 h.</p>
Full article ">Figure 6
<p>Diurnal spatial distribution of the number of flashes at 3-h time intervals in boxes of 0.25 × 0.25 degrees for each season: (<b>a</b>)—winter (DFJ), (<b>b</b>)—spring (MAM), (<b>c</b>)—summer (JJA), (<b>d</b>)—autumn (SON) of the 10 years (Marth 2005–February 2015). The color scale represents the number of flashes; note that the scale is different in each season. Local time is UTC + 2 h or UTC + 3 h.</p>
Full article ">Figure 7
<p>Spatial distribution of number of recorded flashes in boxes of 0.25° × 0.25° over Bulgaria and the Black Sea for September (2005–2014).</p>
Full article ">Figure 8
<p>Mean values of sea surface temperature (SST) for September (2005–2014).</p>
Full article ">Figure 9
<p>Box and Whisker plot of the sea surface temperature (SST) for the cases with and without flashes for all four investigated time-intervals from September months. (median-blue line; 25th–75th percentile, blue box; 10th–90th percentile). LT = UTC + 3 h.</p>
Full article ">Figure 10
<p>Box and Whisker plot of the flash frequency (number of lightning per case) as a function of sea surface temperature (SST) during nighttime intervals ((1800–2100) UTC; (0000–0300) UTC) and daytime intervals ((0600–0900) UTC; (1200–1500) UTC). (trend line of: mean—red line, median—yellow line; 25th–75th percentile, blue box; 10th–90th percentile, whisker; the value in blue box is number of cloud cases). LT = UTC + 3 h. The analysis includes the September months of the period 2005–2014.</p>
Full article ">Figure 11
<p>Spatial–diurnal distribution of flashes at 3-h time intervals: nighttime (1800–2100) UTC, (0000–0300) UTC and daytime (0600–0900) UTC, (1200–1500) UTC for the September months (2005–2014). LT = UTC + 3 h.</p>
Full article ">
20 pages, 10975 KiB  
Article
Numerical Weather Prediction of Sea Surface Temperature in South China Sea Using Attention-Based Context Fusion Network
by Hailun He, Benyun Shi, Yuting Zhu, Liu Feng, Conghui Ge, Qi Tan, Yue Peng, Yang Liu, Zheng Ling and Shuang Li
Remote Sens. 2024, 16(20), 3793; https://doi.org/10.3390/rs16203793 - 12 Oct 2024
Viewed by 686
Abstract
Numerical weather prediction of sea surface temperature (SST) is crucial for regional operational forecasts. Deep learning offers an alternative approach to traditional numerical general circulation models for numerical weather prediction. In our previous work, we developed a sophisticated deep learning model known as [...] Read more.
Numerical weather prediction of sea surface temperature (SST) is crucial for regional operational forecasts. Deep learning offers an alternative approach to traditional numerical general circulation models for numerical weather prediction. In our previous work, we developed a sophisticated deep learning model known as the Attention-based Context Fusion Network (ACFN). This model integrates an attention mechanism with a convolutional neural network framework. In this study, we applied the ACFN model to the South China Sea to evaluate its performance in predicting SST. The results indicate that for a 1-day lead time, the ACFN model achieves a Mean Absolute Error of 0.215 °C and a coefficient of determination (R2) of 0.972. In addition, in situ buoy data were utilized to validate the forecast results. The Mean Absolute Error for forecasts using these data increased to 0.500 °C for a 1-day lead time, with a corresponding R2 of 0.590. Comparative analyses show that the ACFN model surpasses traditional models such as ConvLSTM and PredRNN in terms of accuracy and reliability. Full article
Show Figures

Figure 1

Figure 1
<p>Research area: (<b>a</b>) Mean and (<b>b</b>) variation (standard deviation; STD) of Sea Surface Temperature (SST). In (<b>a</b>), isobaths of −2000 and −4000 m are presented by thin and thick gray lines, respectively. Fixed buoys are denoted by magenta triangles.</p>
Full article ">Figure 2
<p>Flowchart of the present work. DL: deep learning.</p>
Full article ">Figure 3
<p>Example of SST forecast. The now-time (<math display="inline"><semantics> <msub> <mi>D</mi> <mn>0</mn> </msub> </semantics></math>) is 2014/01/02. Lead time spans from 0 to 10 days. The <b>left</b> panels are ground truth (OISST). The <b>right</b> three panels are model output, where the forecasting models include ACFN, PredRNN, and ConvLSTM. The isothermal lines of 23 °C, 25 °C, and 28 °C are also shown as solid black lines.</p>
Full article ">Figure 4
<p>Time series of spatial averaged SST forecast in 2014. <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN, and <b>right</b> panels: ConvLSTM. Lead time spans from 1 to 10 days. The spatial averaged SST is performed within the SCS only. Red lines are ground truth, and blue lines are forecasting results.</p>
Full article ">Figure 5
<p>Spatial distribution of MAE (units: °C). <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN, and <b>right</b> panels: ConvLSTM. Lead time spans from 1 to 10 days.</p>
Full article ">Figure 6
<p>Spatial distribution of <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math>. <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN, and <b>right</b> panels: ConvLSTM. Lead time spans from 1 to 10 days. Results below 0.4 are displayed as blank areas.</p>
Full article ">Figure 7
<p>Bar chart of overall regional evaluation metrics. <b>Left</b> panels: MAE (units: °C) and <b>right</b> panels: <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math>. Lead time spans from 1 to 10 days. The spatial averaged SST is performed within the SCS only.</p>
Full article ">Figure 8
<p>Time series of in situ buoy-based SST (units: °C; Buoy 4). Red lines are in situ observations (same in all the panels), and blue lines are forecasting results. <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN, and <b>right</b> panels: ConvLSTM. Lead time spans from 0 to 10 days.</p>
Full article ">Figure 9
<p>Bar chart of forecasting metrics for in situ buoy-based SST. <b>Left</b> panels: MAE (units: °C) and <b>right</b> panels: <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math>. Lead time spans from 1 to 10 days. Red line is the evaluation of ground truth (OISST).</p>
Full article ">Figure 10
<p>Effective lead time (<math display="inline"><semantics> <msub> <mi>T</mi> <mi>e</mi> </msub> </semantics></math>), which satisfies (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>≥</mo> </mrow> </semantics></math> 0.95 and (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mi>R</mi> <mn>2</mn> </msup> <mo>≥</mo> </mrow> </semantics></math> 0.95 and MAE ≤ 0.4 °C. Isobaths of −2000 and −4000 m are presented by thin and thick gray lines, respectively.</p>
Full article ">Figure A1
<p>Time line of buoy-based SST observations. The locations of buoys are shown in <a href="#remotesensing-16-03793-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure A2
<p>Spatial distribution of RMSE (units: °C). <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN and <b>right</b> panels: ConvLSTM. Lead time spans from 1 to 10 days.</p>
Full article ">Figure A3
<p>Spatial distribution of CC. <b>Left</b> panels: ACFN, <b>middle</b> panels: PredRNN and <b>right</b> panels: ConvLSTM. Lead time spans from 1 to 10 days.</p>
Full article ">
14 pages, 274 KiB  
Article
The Relation between Parameters of Physical Performance and Depression in Consecutive Hospitalized Geriatric Patients with Heart Failure
by Malgorzata Kupisz-Urbańska, Urszula Religioni, Wiktoria Niegowska, Julia Szydlik, Piotr Czapski, Siamala Sinnadurai, Katarzyna Januszewska, Ada Sawicka, Agnieszka Drab, Jarosław Pinkas and Piotr Jankowski
Nutrients 2024, 16(19), 3392; https://doi.org/10.3390/nu16193392 - 6 Oct 2024
Viewed by 1218
Abstract
Background: In the geriatric population, the risk of cardiometabolic diseases is strongly influenced by comorbidities. The aim of the study was to estimate the prevalence of depression among hospitalized patients with heart failure (HF) and to assess the relation between physical performance and [...] Read more.
Background: In the geriatric population, the risk of cardiometabolic diseases is strongly influenced by comorbidities. The aim of the study was to estimate the prevalence of depression among hospitalized patients with heart failure (HF) and to assess the relation between physical performance and depression in this population. Methods: We included consecutive hospitalized patients with HF aged >65 years. The depression symptoms were evaluated using the Geriatric Depression Scale (GDS), the physical performance was assessed using the grip strength measurements, the Back Scratch Test, the Timed Up and Go Test (TUGT), the Five Times Sit to Stand Test (5 × SST), and the 6 min walk test. Results: We included 206 patients (134 females and 72 males, median age 82 years (77–86) years). Altogether, 33% of participants had signs of depression. The association was found between depression severity and economic status (p = 0.001), stressful events (p = 0.005), self-reported general health status (p = 0.001), and heart failure severity assessed by the New York Heart Association class (NYHA), p = 0.003. The Back Scratch Test, the TUGT, and the 5xSST were associated with depression severity in a univariable regression analysis (β coefficient 0.04 [95% CI 0.00–0.08], 0.20 [95% CI 0.12–0.27], 0.18 [95% CI 0.07–0.27], respectively); however, when adjusted for co-factors, the TUGT and the 5xSST (0.17 [95% CI 0.08–0.26] and 0.14 [95% CI 0.02–0.26], respectively) were significantly related to the GDS score. Grip strength and the 6 min walk test were not related to the GDS score in the univariable nor multivariable analysis. These findings were confirmed in the logistic analyses. Conclusions: Our study indicated a high incidence of depression among elderly hospitalized patients with heart failure. Depression severity in older patients with HF is related to physical performance decline as assessed by the Timed Up and Go Test and the Five Times Sit to Stand Test. Grip strength and the 6 min walk test are not related to the GDS score in this population. Full article
17 pages, 11892 KiB  
Article
The Mesoscale SST–Wind Coupling Characteristics in the Yellow Sea and East China Sea Based on Satellite Data and Their Feedback Effects on the Ocean
by Chaoran Cui and Lingjing Xu
J. Mar. Sci. Eng. 2024, 12(10), 1743; https://doi.org/10.3390/jmse12101743 - 3 Oct 2024
Viewed by 651
Abstract
The mesoscale interaction between sea surface temperature (SST) and wind is a crucial factor influencing oceanic and atmospheric conditions. To investigate the mesoscale coupling characteristics of the Yellow Sea and East China Sea, we applied a locally weighted regression filtering method to extract [...] Read more.
The mesoscale interaction between sea surface temperature (SST) and wind is a crucial factor influencing oceanic and atmospheric conditions. To investigate the mesoscale coupling characteristics of the Yellow Sea and East China Sea, we applied a locally weighted regression filtering method to extract mesoscale signals from Quik-SCAT wind field data and AMSR-E SST data and found that the mesoscale coupling intensity is stronger in the Yellow Sea during the spring and winter seasons. We calculated the mesoscale coupling coefficient to be approximately 0.009 N·m−2/°C. Subsequently, the Tikhonov regularization method was used to establish a mesoscale empirical coupling model, and the feedback effect of mesoscale coupling on the ocean was studied. The results show that the mesoscale SST–wind field coupling can lead to the enhancement of upwelling in the offshore area of the East China Sea, a decrease in the upper ocean temperature, and an increase in the eddy kinetic energy in the Yellow Sea. Diagnostic analyses suggested that mesoscale coupling-induced variations in horizontal advection and surface heat flux contribute most to the variation in SST. Moreover, the increase in the wind energy input to the eddy is the main factor explaining the increase in the eddy kinetic energy. Full article
(This article belongs to the Special Issue Air-Sea Interaction and Marine Dynamics)
Show Figures

Figure 1

Figure 1
<p>The probability distributions of the mesoscale magnitude of SST perturbations as a function of the different half-span parameters.</p>
Full article ">Figure 2
<p>The flow chart of mesoscale wind field calculation in MESO-E.</p>
Full article ">Figure 3
<p>Spatially high-pass filtered WS<sub>meso</sub> (contours) and SST<sub>meso</sub> (colors) in the different months in 2006. The contour interval is 0.003 N·m<sup>−2</sup>. The zero contours are not included.</p>
Full article ">Figure 4
<p>Spatially high-pass filtered (upper panel) Div(WS<sub>meso</sub>) (contours) and ∇<sub>down</sub> SST<sub>meso</sub> (colors), and (lower panel) Curl(WS<sub>meso</sub>) (contours) and ∇<sub>cross</sub> SST<sub>meso</sub> (colors) in (<b>a</b>,<b>e</b>) February, (<b>b</b>,<b>f</b>) June, (<b>c</b>,<b>g</b>) August, and (<b>d</b>,<b>h</b>) December 2006. The contour interval is 0.3 N·m<sup>−2</sup> per 10,000 km. The zero contours are not included.</p>
Full article ">Figure 5
<p>Scatterplots of the spatially high-pass filtered Quik-SCAT Div(WS<sub>meso</sub>) and Curl(WS<sub>meso</sub>) binned by ranges of AMSR-E ∇<sub>down</sub> SST<sub>meso</sub> and ∇<sub>cross</sub> SST<sub>meso</sub> perturbations. The coupling coefficient is denoted as S. Points and error bars represent the mean and standard deviation in each bin, respectively.</p>
Full article ">Figure 6
<p>The monthly variations in the coupling coefficient (N·m<sup>−2</sup>/(°C·100 km)) between (<b>a</b>) Div(WS<sub>meso</sub>) and ∇<sub>down</sub> SST<sub>meso</sub> and (<b>b</b>) Curl(WS<sub>meso</sub>) and ∇<sub>cross</sub> SST<sub>meso</sub>.</p>
Full article ">Figure 7
<p>The SST<sub>meso</sub> (colors) and WS<sub>meso</sub> (contours) obtained from (<b>a</b>) observation, (<b>b</b>) MESO-E, and (<b>c</b>) CONTROL-E in summer; the SST<sub>meso</sub> (colors) and WS<sub>meso</sub> (contours) obtained from (<b>d</b>) observation, (<b>e</b>) MESO-E, and (<b>f</b>) CONTROL-E in winter. The observations are from AMSR-E and Quik-SCAT data in 2006; the simulated results are from 10-year averaged outputs of MESO-E and CONTROL-E. The contour interval is 0.006 N·m<sup>−2</sup>. The zero contours are omitted.</p>
Full article ">Figure 8
<p>The (<b>a</b>) 6-year averaged coupling coefficient between Quik-SCAT WS<sub>meso</sub> and AMSR-E SST<sub>meso</sub>, (<b>b</b>) 10-year averaged coupling coefficient between WS<sub>meso</sub> and SST<sub>meso</sub> from the MESO-E output, and (<b>c</b>) 10-year averaged coupling coefficient between WS<sub>meso</sub> and SST<sub>meso</sub> from the CONTROL-E output. Points and error bars represent the mean and standard deviation in each bin, respectively.</p>
Full article ">Figure 9
<p>The 10-year averaged differences (MESO-E minus CONTROL-E) in (<b>a</b>) sea temperature, (<b>b</b>) surface heat flux, (<b>c</b>) horizontal advection, and (<b>d</b>) vertical diffusion in the upper 50 m. The units are °C in (<b>a</b>) and °C/month in (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 10
<p>The 10-year average (<b>a</b>–<b>c</b>) zonal and (<b>d</b>–<b>f</b>) meridional current differences (m/s) in winter between the SODA3.4.2 2011–2020 data and CONTROL-E (SODA minus CONTROL-E, <b>left panel</b>); between MESO-E and CONTROL-E (MESO-E minus CONTROL-E, <b>middle panel</b>), and between MESO-E and SODA (MESO-E minus SODA, <b>right panel</b>). The zonal and meridional current was calculated by averaging vertically up to a depth of 50 m.</p>
Full article ">Figure 11
<p>Differences in the 10-year average (left panel) Curl(WS<sub>meso</sub>) (1 × 10<sup>−6</sup> N/m<sup>3</sup>) and (right panel) vertical current (1 × 10<sup>−7</sup> m/s) in (<b>a</b>,<b>b</b>) winter and (<b>c</b>,<b>d</b>) summer between MESO-E and CONTROL-E (MESO-E minus CONTROL-E). The vertical current was calculated by averaging vertically up to a depth of 50 m.</p>
Full article ">Figure 12
<p>The 10-year averaged difference in (<b>a</b>) EKE, (<b>b</b>) eddy wind work, and (<b>c</b>) baroclinic conversion from eddy available potential energy to EKE. (<b>d</b>) Conversion between mean kinetic energy and EKE between MESO-E and CONTROL-E (MESO-E minus CONTROL-E). The units are cm<sup>3</sup>/s<sup>3</sup> in (<b>a,b,c</b>) and 1 × 10<sup>−2</sup> cm<sup>3</sup>/s<sup>3</sup> in (<b>d</b>).</p>
Full article ">
19 pages, 5637 KiB  
Article
A Zonal Detached Eddy Simulation of the Trailing Edge Stall Process of a LS0417 Airfoil
by Wenbo Shi, Heng Zhang and Yuanxiang Li
Aerospace 2024, 11(9), 731; https://doi.org/10.3390/aerospace11090731 - 6 Sep 2024
Viewed by 662
Abstract
A Zonal Detached Eddy Simulation (ZDES) based on the SST turbulence model is implemented to the numerical investigation of the trailing edge stall of a LS-0417 airfoil, which includes multiple DES modes for different classifications of flow separation and adopts the subgrid scale [...] Read more.
A Zonal Detached Eddy Simulation (ZDES) based on the SST turbulence model is implemented to the numerical investigation of the trailing edge stall of a LS-0417 airfoil, which includes multiple DES modes for different classifications of flow separation and adopts the subgrid scale definition of Δω. The entire stall process under a series of AOA is simulated according to the experiment condition. The performance of URANS and ZDES in the prediction of the stall flow field are compared. The results reveal that the stall point obtained through ZDES is consistent with the experiment; the deviation of the predicted maximum lift coefficient from the measured result is only 0.8%, while the maximum lift is overpredicted by both RANS and URANS. The high frequency fluctuations are observed in the time history of the lift in ZDES result during stall. With the increase in the AOA, a mild development of separation and a gradual decrease in leading edge peak suction are manifested in the ZDES result. The alternate shedding of shear layers and the interference between the leading edge and trailing edge vortices are illustrated through ZDES near the stall point; the corresponding turbulent fluctuations with high intensity are captured in the separation region, which indicates the essential difference in the prediction of stall process between URANS and ZDES. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>The definition of <math display="inline"><semantics> <mrow> <msub> <mo>Δ</mo> <mi>ω</mi> </msub> </mrow> </semantics></math> of a grid cell.</p>
Full article ">Figure 2
<p>The topology and section of the computational grid of LS-0417.</p>
Full article ">Figure 3
<p>The definition of calculation region of ZDES.</p>
Full article ">Figure 4
<p>The convergence history of the lift coefficient with different methods.</p>
Full article ">Figure 5
<p>The convergence history of lift coefficient with different time steps.</p>
Full article ">Figure 6
<p>Comparison of lift prediction results with different methods.</p>
Full article ">Figure 7
<p>The convergence history of lift coefficient obtained with URANS and ZDES at different AOAs.</p>
Full article ">Figure 7 Cont.
<p>The convergence history of lift coefficient obtained with URANS and ZDES at different AOAs.</p>
Full article ">Figure 8
<p>Comparison of time-averaged streamlines between URANS and ZDES.</p>
Full article ">Figure 8 Cont.
<p>Comparison of time-averaged streamlines between URANS and ZDES.</p>
Full article ">Figure 9
<p>Comparison of time-averaged pressure distribution between URANS and ZDES.</p>
Full article ">Figure 10
<p>Comparison of instantaneous vorticity between URANS and ZDES.</p>
Full article ">Figure 10 Cont.
<p>Comparison of instantaneous vorticity between URANS and ZDES.</p>
Full article ">Figure 11
<p>Comparison of time-averaged turbulent fluctuation between URANS and ZDES, α = 20°.</p>
Full article ">Figure 12
<p>Comparison of iso-surface of Q-criterion between URANS and ZDES.</p>
Full article ">
26 pages, 5887 KiB  
Article
Computational Fluid Dynamics Analyses on How Aerodynamic Rule Changes Impact the Performance of a NASCAR Xfinity Racing Series Racecar
by Mesbah Uddin and Nazarii Olkhovskyi
Vehicles 2024, 6(3), 1545-1570; https://doi.org/10.3390/vehicles6030073 - 31 Aug 2024
Viewed by 835
Abstract
The Xfinity Racing Series is an American stock car racing series organized by NASCAR. For the 2017 racing season, NASCAR introduced new regulations with the objective of creating a level playing field by reducing aerodynamic influence on vehicle performance. In this context, the [...] Read more.
The Xfinity Racing Series is an American stock car racing series organized by NASCAR. For the 2017 racing season, NASCAR introduced new regulations with the objective of creating a level playing field by reducing aerodynamic influence on vehicle performance. In this context, the primary objective of this work is to explore the differences in the aerodynamic performance between the 2016 and 2017 Toyota Camry Xfinity racecars using only open-source Computational Fluid Dynamics (CFD) and CAE tools. During the CFD validation process, it was observed that none of the standard turbulence models, with default turbulence model closure coefficients, were able to provide racecar aerodynamic characteristics predictions with acceptable accuracy compared to experiments. This necessitated a fine-tuning of the closure coefficient numeric values. This work also demonstrates that it is possible to generate CFD predictions that are highly correlated with experimental measurements by modifying the closure coefficients of the standard kω SST turbulence model. Full article
Show Figures

Figure 1

Figure 1
<p>Toyota Xfinity Series racecars. (<b>Top</b>): 2016 car; (<b>Middle</b>): 2017 car; (<b>Bottom</b>): 2016 and 2017 rear spoiler differences.</p>
Full article ">Figure 2
<p>Comparison of the 2016 and 2017 season splitter configurations, with the 2017 configurations highlighted in red. The 2017 car features a smaller splitter surface (<b>Top</b>) and a lower splitter gap (<b>Bottom</b>) compared to the 2016 model.</p>
Full article ">Figure 3
<p>Computational domain; arrow denotes flow direction going from left to right.</p>
Full article ">Figure 4
<p>Volume mesh and different mesh refinement levels; slices through <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> <mi>z</mi> </mrow> </semantics></math> (<b>Top</b>), <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> (<b>Middle</b>), and <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> <mi>z</mi> </mrow> </semantics></math> (<b>Bottom</b>) planes.</p>
Full article ">Figure 5
<p>Zoomed-in view of the volume mesh showing refinement levels at various distances forming the surface around the splitter region.</p>
Full article ">Figure 6
<p>Oil-flow lines superimposed on wall shear stress contour, as obtained from CFD simulations using default (<b>Top</b>) and modified (<b>Bottom</b>) values of <math display="inline"><semantics> <msub> <mi>a</mi> <mn>1</mn> </msub> </semantics></math>.</p>
Full article ">Figure 7
<p>Aerodynamic characteristic differences between 2016 and 2017 cars from CFD simulations and wind tunnel tests.</p>
Full article ">Figure 8
<p>The contributions of the different components of the racecar to <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>C</mi> <mi>D</mi> </msub> </mrow> </semantics></math> (<b>Top</b>) and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>C</mi> <mi>L</mi> </msub> </mrow> </semantics></math> (<b>Bottom</b>) between the 2016 and 2017 configurations.</p>
Full article ">Figure 9
<p>Static pressure coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) distribution on the upper surface of the 2016 (<b>Top</b>) and 2017 (<b>Bottom</b>) cars.</p>
Full article ">Figure 10
<p>Static pressure coefficient (<math display="inline"><semantics> <msub> <mi>C</mi> <mi>P</mi> </msub> </semantics></math>) distribution on the underbody surface of the 2016 (<b>Top</b>) and 2017 (<b>Bottom</b>) cars.</p>
Full article ">Figure 11
<p>Locations of the pressure probes. Please note that in the bottom figure, the 2017 spoiler is overlaid on the 2016 spoiler in pink to better illustrate the locations of the pressure probes.</p>
Full article ">Figure 12
<p>Pressure probe data.</p>
Full article ">Figure 13
<p>Velocity distribution under the car at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mi>h</mi> <mo>/</mo> <mn>2</mn> </mrow> </semantics></math>, where <span class="html-italic">h</span> is the splitter-gap, for the 2016 (<b>Top</b>) and 2017 (<b>Bottom</b>) cars.</p>
Full article ">Figure 14
<p>Pole speed for different race tracks as recorded during 2016 and 2017 race seasons. Source: Jayski’s NASCAR Silly Season Site; <a href="https://www.jayski.com/xfinity-series/2016-nascar-xfinityseries-race-results/" target="_blank">https://www.jayski.com/xfinity-series/2016-nascar-xfinityseries-race-results/</a> for 2016 car and <a href="https://www.jayski.com/xfinity-series/2017-nascar-xfinityseries-race-results/" target="_blank">https://www.jayski.com/xfinity-series/2017-nascar-xfinityseries-race-results/</a> for 2017 car (accessed on 31 August 2024).</p>
Full article ">
Back to TopTop