Nothing Special   »   [go: up one dir, main page]

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

Search Results (108)

Search Parameters:
Keywords = transport of intensity equation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 20230 KiB  
Article
Understanding Emission Trends, Regional Distribution Differences, and Synergistic Emission Effects in the Transportation Sector in Terms of Social Factors and Energy Consumption
by Yu Zhao and Prasanna Divigalpitiya
Sustainability 2024, 16(24), 10971; https://doi.org/10.3390/su162410971 - 13 Dec 2024
Viewed by 437
Abstract
China’s transportation sector plays a significant role in reducing carbon dioxide (CO2) and air pollution. Previous studies have predominantly utilized scenario analysis to forecast emissions for the next 30 to 50 years based on coefficients from a base year. To elucidate [...] Read more.
China’s transportation sector plays a significant role in reducing carbon dioxide (CO2) and air pollution. Previous studies have predominantly utilized scenario analysis to forecast emissions for the next 30 to 50 years based on coefficients from a base year. To elucidate the current state of gas emissions in the transportation sector, this study employed panel data for 10 types of gas emissions from 2001 to 2020, analyzing their emission characteristics, tendencies, and synergistic effects. Utilizing the Kaya equation and the logarithmic mean division index (LMDI) decomposition method, we developed a model of pollutant emissions that considers the synergistic effects, pollution emission intensity, energy mix, energy consumption intensity, and population. The results show that all pollutants in the transportation sector decreased except for NH3 and CO2. There was a synergistic effect between air pollutants and CO2 emissions, but the reduction was not significant. From 2013 to 2020, the transportation sector shifted from a high emission intensity with low synergy to a low emission intensity with high synergy. The results indicate that off-road mobile vehicles, on-road diesel vehicles, and motorcycles became the main source of emissions from transportation in certain provinces, and a key area requiring attention in policy development. Gasoline consumption was identified as the primary contributor to the significant increase in synergistic emission variability in the transportation sector. These results provide policymakers with practical ways to optimize emission reduction pathways. Full article
Show Figures

Figure 1

Figure 1
<p>Analysis structure. Note: this study analyzed 10 types of emissions from four pollution sources. On the left side is the sequence of research steps, while the corresponding research methods are presented on the right side (source: created by the authors).</p>
Full article ">Figure 2
<p>Sources of pollutant emissions in 2020.</p>
Full article ">Figure 3
<p>Sources of pollutants and changes in emissions. Note: The subfigures (<b>a</b>–<b>j</b>) show sources of CO<sub>2</sub>, CO, PM<sub>2.5</sub>, PM<sub>10</sub>, BC, OC, VOC, NOx, SO<sub>2</sub>, NH<sub>3</sub>, separately.</p>
Full article ">Figure 4
<p>Changes relative to the previous year.</p>
Full article ">Figure 5
<p>Emission intensity and synergistic effects in 2013 and 2020. Note: The subfigures (<b>a</b>–<b>f</b>) show synergistic effect of the pollutant with CO<sub>2</sub>. The red dots mark the average of the 30 provinces and cities.</p>
Full article ">Figure 6
<p>Variations in emission origins across four provinces and cities. Note: The subfigures (<b>a</b>–<b>f</b>) show sources of PM<sub>2.5</sub>, PM<sub>10</sub>, NOx, VOC, SO<sub>2</sub>, CO<sub>2</sub>, separately.</p>
Full article ">Figure 7
<p>Theil index with population and gasoline as the base.</p>
Full article ">Figure 8
<p>Province categories.</p>
Full article ">Figure 9
<p>Emission origins of the seven major regions.</p>
Full article ">
19 pages, 1368 KiB  
Article
Determinants of Dockless Bicycle-Sharing Adoption and Usage Intensity for Commuting and Errands: Evidence from Disadvantaged Neighborhoods
by Hongyu Wang, Yu Dong and Liang Ma
Land 2024, 13(12), 2055; https://doi.org/10.3390/land13122055 - 30 Nov 2024
Viewed by 548
Abstract
The emergence of dockless bicycle sharing has transformed urban transportation, particularly in China, by offering a flexible and accessible travel option. However, understanding the factors driving its adoption and usage in disadvantaged neighborhoods is crucial, as these areas often face unique mobility challenges. [...] Read more.
The emergence of dockless bicycle sharing has transformed urban transportation, particularly in China, by offering a flexible and accessible travel option. However, understanding the factors driving its adoption and usage in disadvantaged neighborhoods is crucial, as these areas often face unique mobility challenges. This study explores these determinants, providing a more comprehensive analysis than prior research by focusing specifically on disadvantaged communities. Using survey data from four such neighborhoods in Xi’an, China, we apply Structural Equation Modeling to investigate the factors influencing individuals’ decisions to adopt and intensively use dockless bicycle sharing for commuting and errands. The results reveal key determinants, including psychological factors, demographic characteristics, and spatial and social contexts, and their interaction mechanisms. Attitudes are found to have a substantial impact on bicycle-sharing behavior for both commuting and errands, while social norms and perceived behavioral control (PBC) mainly influence usage for errands. Interestingly, PBC affects adoption but not usage frequency. The findings also highlight that proximity to schools, subways, and neighborhood aesthetics positively correlate with bicycle-sharing adoption for errands, whereas bicycling infrastructure significantly influences usage intensity. However, none of the neighborhood environment factors were found to significantly affect adoption for commuting purposes. These insights are especially valuable for developing targeted strategies to promote bicycle sharing as a sustainable transportation solution in disadvantaged neighborhoods, where improved access can significantly enhance mobility and quality of life. Full article
Show Figures

Figure 1

Figure 1
<p>Conceptual framework.</p>
Full article ">Figure 2
<p>Different bicycle-sharing usage behaviors of the same respondents. Notes: True = bicycle sharing is the primary mode; False = bicycle sharing is not the primary mode; Always = “more than twice a week”; Usually = “once a week”; Often = “about once every two weeks”; Sometimes = “once or twice a month”; Rarely = “less than once a month”.</p>
Full article ">Figure 3
<p>Summary of model results.</p>
Full article ">
28 pages, 26490 KiB  
Article
Vertical Accelerations and Convection Initiation in an Extreme Precipitation Event in the Western Arid Areas of Southern Xinjiang
by Na Li, Lingkun Ran, Daoyong Yang, Baofeng Jiao, Cha Yang, Wenhao Hu, Qilong Sun and Peng Tang
Atmosphere 2024, 15(12), 1406; https://doi.org/10.3390/atmos15121406 - 22 Nov 2024
Viewed by 372
Abstract
A simulation of an extreme precipitation event in southern Xinjiang, which is the driest area in China, seizes the whole initiation process of the intense convective cell responsible for the high hourly rainfall amount. Considering the inner connection between convection and vertical motions, [...] Read more.
A simulation of an extreme precipitation event in southern Xinjiang, which is the driest area in China, seizes the whole initiation process of the intense convective cell responsible for the high hourly rainfall amount. Considering the inner connection between convection and vertical motions, the characteristics and mechanisms of the vertical accelerations during this initial development of the deep convection are studied. It is shown that three key accelerations are responsible for the development from the nascent cumuli to a precipitating deep cumulonimbus, including sub-cloud boundary-layer acceleration, in-cloud deceleration, and cloud-top acceleration. By analyzing the right-hand terms of the vertical velocity equation in the framework of the WRF model, together with a diagnosed relation of perturbation pressure to perturbation potential temperature, perturbation-specific volume (or density), and moisture, the physical processes associated with the corresponding accelerations are revealed. It is found that sub-cloud acceleration is associated with three-dimensional divergence, indicating that the amount of upward transported air must be larger than that of horizontally convergent air. This is favorable for the persistent accumulation of water vapor into the accelerated area. In-cloud deceleration is caused by the intrusion or entrainment of mid-level cold air, which cools down the developing cloud and delays the deep convection formation. Cloud-top acceleration is responsible for the rapid upward extension of the cloud top, which is highly correlated with the convergence and upward transport of moisture. Full article
(This article belongs to the Section Meteorology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Accumulated 24 h precipitation on 15 June 2021. (<b>b</b>) Bar chart of hourly precipitation over several automatic observation stations in southern Xinjiang.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>i</b>) Observed composite radar reflectivity from the C-band radar in Hoton in southern Xinjiang. The stations indicated by the black stars are, respectively, Sampoulu (“1”), Lop (“2”), Hoton (“3”), and Moyu (“4”). The black box and the red circle indicate the focused convective areas.</p>
Full article ">Figure 3
<p>The weather pattern at 1100 UTC on 15 June 2021. (<b>a</b>) Geopotential height (black contours, unit: 1 gpm), wind speed (color shaded, unit: m s<sup>−1</sup>) and horizontal divergence regions (red contours, unit: 10<sup>−5</sup> s<sup>−1</sup>, 4 interval) at 200 hPa; (<b>b</b>) geopotential height (black contours, unit: 10 gpm) at 500 hPa, and wind speed (color shaded) and water vapor flux (red arrows, unit:) at 800 hPa. The symbol “H” means the high-pressure system and “L” means the low-pressure system, which is denoted by geopotential height in the figures. The black dotted short lines are trough regions. The black boxes indicate the focused precipitation region. The yellow arrow box indicates the flow from the east–west-oriented shallow trough, the pink arrow box indicates the flow from the north–south-oriented shallow trough and the green arrow indicates the flow from the east–west-oriented shallow trough over the mountain. The black thick arrows indicate the main directions of the moisture transportation.</p>
Full article ">Figure 4
<p>Simulated maximum radar reflectivities at (<b>a</b>) 0700 UTC, (<b>b</b>) 0800 UTC, (<b>c</b>) 0900 UTC, (<b>d</b>) 1000 UTC, (<b>e</b>) 1100 UTC, (<b>f</b>) 1200 UTC on 15 June 2021. Boxes enclosed the convective cell that the paper concerned. (<b>g</b>) The bar chart is the 10-min precipitation during this period.</p>
Full article ">Figure 4 Cont.
<p>Simulated maximum radar reflectivities at (<b>a</b>) 0700 UTC, (<b>b</b>) 0800 UTC, (<b>c</b>) 0900 UTC, (<b>d</b>) 1000 UTC, (<b>e</b>) 1100 UTC, (<b>f</b>) 1200 UTC on 15 June 2021. Boxes enclosed the convective cell that the paper concerned. (<b>g</b>) The bar chart is the 10-min precipitation during this period.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Maximum radar reflectivity (shaded, unit: dBz); (<b>c</b>,<b>d</b>) radar reflectivity (shaded, unit: dBz) and vertical velocity (contour, unit: m s<sup>−1</sup>); and (<b>e</b>,<b>f</b>) vertical accelerations denoted by Net_WAη (shaded, unit: 10<sup>−3</sup> s<sup>−1</sup>) and sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 black contours, unit: g/kg) at 0750 UTC along 78.5° E (left column) and 0820 UTC along 78.6° E (right column). The purple contour in (<b>f</b>) is 35 dBz radar reflectivity. The thick black line is 0 °C isotherm contour.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) Maximum radar reflectivity (shaded, unit: dBz); (<b>d</b>–<b>f</b>) radar reflectivity (shaded, unit: dBz) and vertical velocity (contour, unit: m s<sup>−1</sup>); and (<b>g</b>–<b>i</b>) vertical accelerations denoted by Net_WAη (shaded, unit: 10<sup>−3</sup> s<sup>−1</sup>) and the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) at 0910 UTC (left column), 0920 (middle column) and 0930 UTC (right column) along 78.8° E. The purple contour in (<b>g</b>–<b>i</b>) is 35 dBz radar reflectivity. The thick black line is the 0 °C isotherm contour. The green lines in (<b>d</b>–<b>f</b>) indicate 1 h precipitation (unit: mm) with magnitude on the right side of the y-axis.</p>
Full article ">Figure 7
<p>(<b>a</b>) PGF_WRF, (<b>b</b>) B_WRF and (<b>c</b>) Net_WAη at 0710 UTC along 78.5° N (contours, unit: 10<sup>−3</sup> s<sup>−1</sup>); (<b>d</b>–<b>f</b>) same as (<b>a</b>–<b>c</b>) but for 0750 UTC; (<b>g</b>) PGF_WRF difference, (<b>h</b>) B_WRF difference, and (<b>i</b>) Net_WAη difference between 0750 UTC and 0710 UTC along 78.5° N. The shaded areas are radar reflectivities. The red contours are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud. The “difference” means the fields at 0750 UTC minus those at 0710 UTC. The colored boxes indicate the areas with evident differences between 0710 and 0750 UTC. The areas enclosed by the orange box, blue box and green box are, respectively, the in-cloud area, the sub-cloud area and the boundary-layer area in front of the convection.</p>
Full article ">Figure 8
<p>(<b>a</b>) Difference of perturbation pressure between 0710 UTC and 0750 UTC (contours, unit: hPa) along 78.5° N; (<b>b</b>) same as (<b>a</b>) but for <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>θ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>; (<b>c</b>) same as (<b>a</b>) but for <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>ρ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>; (<b>d</b>) same as (<b>a</b>) but for <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>ρ</mi> <mo>′</mo> </msubsup> <mo>+</mo> <msubsup> <mi>p</mi> <mi>θ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>; (<b>e</b>) same as (<b>a</b>) but for <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>m</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>. The shaded areas are radar reflectivities. The red contours are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud area. The “difference” means the fields at 0750 UTC minus those at 0710 UTC. The red thick arrows indicate the direction of PGF due to the change in the corresponding perturbation pressure.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Water vapor mixing ratio (shaded and black contours, unit: g/kg); (<b>c</b>,<b>d</b>) magnitude of water vapor flux (shaded) and wind vectors of (v, w) (unit: m/s); (<b>e</b>,<b>f</b>) horizontal wind speed; (<b>g</b>,<b>h</b>) 2D horizontal divergence (shaded, unit: 10<sup>−4</sup> s<sup>−1</sup>) and wind vectors of (v, w) (unit: m/s); and (<b>i</b>,<b>j</b>) 3D horizontal divergence (shaded, unit: 10<sup>−4</sup> s<sup>−1</sup>) and wind vectors of (v, w) (unit: m/s) in the section at 0710 UTC (left column) and 0750 UTC (right column) along 78.5° N. The red contours are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud. The blue line in (<b>e</b>,<b>f</b>) is the free convection level, and the orange line is the condensation level. The green line in (<b>j</b>) is the minus of the mass in column at 0750 and 0710 UTC (<math display="inline"><semantics> <mrow> <msub> <mrow> <mfenced close="|" open=""> <mrow> <msubsup> <mi>μ</mi> <mi>d</mi> <mo>′</mo> </msubsup> </mrow> </mfenced> </mrow> <mrow> <mn>0750</mn> <mi>U</mi> <mi>T</mi> <mi>C</mi> </mrow> </msub> <mo>−</mo> <msub> <mrow> <mfenced close="|" open=""> <mrow> <msubsup> <mi>μ</mi> <mi>d</mi> <mo>′</mo> </msubsup> </mrow> </mfenced> </mrow> <mrow> <mn>0710</mn> <mi>U</mi> <mi>T</mi> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 10
<p>(<b>a</b>) PGF_WRF, (<b>b</b>) B_WRF and (<b>c</b>) their sum for Net_WAη at 0820 UTC (contours, unit: 10<sup>−3</sup> s<sup>−1</sup>) along 78.6° N. The shaded areas are radar reflectivities. The red contours are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud.</p>
Full article ">Figure 11
<p>(<b>a</b>) Perturbation pressure, (<b>b</b>) <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>θ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>ρ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>ρ</mi> <mo>′</mo> </msubsup> <mo>+</mo> <msubsup> <mi>p</mi> <mi>θ</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math>; (<b>e</b>) <math display="inline"><semantics> <mrow> <msubsup> <mi>p</mi> <mi>m</mi> <mo>′</mo> </msubsup> </mrow> </semantics></math> at 0820 UTC along 78.6° N (contours, unit: hPa). The shaded areas are radar reflectivities. The red contours are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud.</p>
Full article ">Figure 12
<p>(<b>a</b>) Perturbation temperature at 0750 UTC (shaded and contours, unit: K), (<b>b</b>) perturbation temperature at 0820 UTC (shaded and contours, unit: K), (<b>c</b>) diabatic heating and cooling associated with the microphysical process at 0820 UTC (shaded, unit: 10<sup>−3</sup> K s<sup>−1</sup>), (<b>d</b>) the 2D divergence field at 0820 UTC (shaded, unit: 10<sup>−4</sup> s<sup>−1</sup>) along 78.6° N. The red contours in (<b>a</b>,<b>b</b>) and black contours in (<b>c</b>,<b>d</b>) are the sum of mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud. The vectors are the wind vectors of (v, w) (unit: m/s) in the current section.</p>
Full article ">Figure 13
<p>(<b>a</b>) The 2D divergence fields (shaded, unit: 10<sup>−4</sup> s<sup>−1</sup>) and horizontal wind vectors (arrows, unit: m/s) at 7 km height at 0750 UTC, (<b>b</b>) 0820 UTC, (<b>c</b>) 0910 UTC and (<b>d</b>) 0930 UTC. The purple contours are 35 dBz radar reflectivity.</p>
Full article ">Figure 14
<p>Same as <a href="#atmosphere-15-01406-f007" class="html-fig">Figure 7</a> except for the time of 0850 UTC and 0910 UTC. The “difference” means the fields at 0910 UTC minus those at 0850 UTC. The area enclosed by the thick black line is the positive Net_WAη area, while that enclosed by the rose red line is the negative Net_WAη area. “PGF” is the pressure gradient force, and “B” is buoyancy.</p>
Full article ">Figure 15
<p>Same as <a href="#atmosphere-15-01406-f008" class="html-fig">Figure 8</a> except for the time of 0850 UTC and 0910 UTC. The “difference” means the fields at 0910 UTC minus those at 0850 UTC. The red thick arrows indicate the direction of PGF due to the change in the corresponding perturbation pressure.</p>
Full article ">Figure 16
<p>(<b>a</b>,<b>b</b>) Diabatic heating and cooling associated with the microphysical process (shaded, unit: 10<sup>−3</sup> K s<sup>−1</sup>); (<b>c</b>,<b>d</b>) water vapor mixing ratio (shaded and black contours, unit: g/kg) in the section at 0850 UTC (left column) and 0910 UTC (right column) along 78.8° N. The black contours in (<b>a</b>,<b>b</b>) and red contours in (<b>c</b>,<b>d</b>) are the sum of the mixing ratios of ice and cloud particles (0.05, 0.1, 0.2, 0.3, 0.4 values, unit: g/kg) to indicate the cloud. The purple contours are 35 dBz radar reflectivity.</p>
Full article ">Figure 17
<p>Horizontal wind speed (left column, shaded, unit: m/s) and water vapor mixing ratio (right column, shaded and contours, unit: 10<sup>−3</sup> g Kg<sup>−1</sup>) at 2 km height at (<b>a</b>,<b>b</b>) 0610 UTC, (<b>c</b>,<b>d</b>) 0700 UTC, (<b>e</b>,<b>f</b>) 0800 UTC, (<b>g</b>,<b>h</b>) 0850 UTC, and (<b>i</b>,<b>j</b>) 0930 UTC. The white, purple and red contours, respectively, represent the 10, 35 and 50 dBz radar reflectivity. The arrows are wind vectors. The black box indicates the location of the focused convection. The colored tringles indicate the small-scale ridges. The thick arrow in (<b>a</b>,<b>b</b>) indicates the moving direction of the topographic clouds.</p>
Full article ">
19 pages, 8819 KiB  
Article
Experimental Investigation of the Spread and Burning Behaviors of Diesel Spill Fires on a Water Surface
by Jinlong Zhao, Yijia He, Jingwen Xiao, Zilong Su, Hanchao Ma and Xu Zhai
Fire 2024, 7(11), 402; https://doi.org/10.3390/fire7110402 - 1 Nov 2024
Viewed by 674
Abstract
Large quantities of water are often used to extinguish fires when accidental fuel leakage occurs during storage and transportation. This may lead to spill fires on water. Boilover and splash of heavy oil spill fires in particular can pose a serious thermal threat [...] Read more.
Large quantities of water are often used to extinguish fires when accidental fuel leakage occurs during storage and transportation. This may lead to spill fires on water. Boilover and splash of heavy oil spill fires in particular can pose a serious thermal threat to surrounding facilities and personnel. In this work, a series of diesel spill fire experiments were conducted on the surface of water. The results showed that, for the non-ignition cases, the fuel spread velocity was fast at first, then maintained a long period of steady spread, which can be successfully predicted by a developed spread model. During the ignition process, the burning of diesel fuel is divided into four phases. Following a brief quasi-steady burning phase, we observed an expansion of the burning area during the intermittent boilover phase, which was primarily driven by boilover. During the quasi-steady burning phase, the burning rate was lower than that of pool fires, which is attributed to the heat loss between the diesel and water layers. This heat loss also results in a lower flame height than the pool fire, and a dimensionless equation was proposed to eliminate discrepancies. During the intermittent boilover phase, the increase in the burning area was used to characterize the boilover intensity, which was found to be negatively correlated with the number of boilovers. Furthermore, the emergence of the boilover also caused flame radiation to rise rapidly; it was about 19% to 30% higher than that in the quasi-steady burning phase. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of: (<b>a</b>) complete experimental setup; (<b>b</b>) thermocouples arrangement in the water layer.</p>
Full article ">Figure 2
<p>Variation in spread distance of diesel fuel with time for different discharge rates.</p>
Full article ">Figure 3
<p>Variation in spread velocity of diesel fuel with time for different discharge rates.</p>
Full article ">Figure 4
<p>The relationship between the one-dimensional spread distance of diesel fuel and the variation in <math display="inline"><semantics> <mrow> <msup> <mrow> <mo stretchy="false">(</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mo>Δ</mo> <mi mathvariant="normal">g</mi> </mrow> <mrow> <msup> <mi>v</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> </mrow> </mfrac> </mstyle> <mo stretchy="false">)</mo> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>4</mn> </mrow> </msup> <msup> <mrow> <msub> <mi mathvariant="normal">Q</mi> <mi mathvariant="normal">t</mi> </msub> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msup> <msup> <mi mathvariant="normal">t</mi> <mrow> <mn>7</mn> <mo>/</mo> <mn>8</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 5
<p>The burning spread behavior during different phases of test 6.</p>
Full article ">Figure 6
<p>The variation in burning spread distance for diesel during the intermittent boilover burning phase under different discharge rates.</p>
Full article ">Figure 7
<p>The variation in the burning area over time at different discharge rates.</p>
Full article ">Figure 8
<p>Comparison of experimental burning rate with that predicted by the model for pool fires and the modified model for spill fires.</p>
Full article ">Figure 9
<p>The heat transfer mechanism in the liquid layer in spill fire.</p>
Full article ">Figure 10
<p>The temperature variation in the water layer 10 cm from the preheating trench in test 6.</p>
Full article ">Figure 11
<p>Flame height treatment process.</p>
Full article ">Figure 12
<p>Variation in the flame height of diesel during the quasi-steady burning phase in relation to discharge rate.</p>
Full article ">Figure 13
<p>Comparison of experimental flame heights with predicted values from the Heskestad model.</p>
Full article ">Figure 14
<p>The variations in boilover intensity during the intermittent boilover phase of diesel at different discharge rates.</p>
Full article ">Figure 15
<p>The variation in radiative heat flux over time in test 6.</p>
Full article ">Figure 16
<p>The average radiative heat flux during the quasi-steady burning phase and the intermittent boilover phase.</p>
Full article ">
28 pages, 9344 KiB  
Article
Multivariate Structural Vibration Coupling Response of the Self-Propelled Straw Pickup Baler Under Time-Varying Loads
by Bangzhui Wang, Kexin Que, Zhong Tang, Meiyan Sun, Yi Lian and Haoyang Wang
Agriculture 2024, 14(11), 1960; https://doi.org/10.3390/agriculture14111960 - 1 Nov 2024
Viewed by 585
Abstract
The self-propelled straw pickup baler in agricultural work is responsible for collecting and compressing straw to facilitate transportation and storage, while reducing waste and environmental pollution. Like other agricultural equipment, the straw pickup baler is a complex mechanical system. During operation, its excitation [...] Read more.
The self-propelled straw pickup baler in agricultural work is responsible for collecting and compressing straw to facilitate transportation and storage, while reducing waste and environmental pollution. Like other agricultural equipment, the straw pickup baler is a complex mechanical system. During operation, its excitation characteristics under multi-source stimuli and the coupling characteristics of various components are not yet clear. This paper analyzed the excitation mechanics property of each component of the self-propelled straw pickup baler and established balance equations. Based on the balance equations, the coupling characteristics of the structures were studied. Through experiments collecting excitation signals from multiple devices under different operating conditions, the vibration excitation signals of each component were obtained. The experiments revealed that the excitation and coupling signals in the Z direction are particularly evident. Based on experiments, the effective Z-direction vibration signal value on the left front of the chassis exceeds 7 m·s2, while on the right front it increases from 1.995 m·s2 to 7.287 m·s2, indicating the most intense vibration direction. It was also found that, at the driver’s cab, the effective Z-direction vibration signal values at two response points, 11 and 12, both exceed 7 m·s2. The data indicate significant vibrations occur in both the longitudinal and vertical directions. Using the Signal Analyzer module in MATLAB for signal processing, it was found that the prominent filtered signals consist of combustion excitation harmonics and continuous low-frequency vibrations from the compression mechanism. The periodic reciprocating compression motion of the crank-slider mechanism causes sustained impacts on the frame, leading to periodic changes in the vibration amplitude of the chassis. Thus, the vibration reduction of the compression mechanism’s periodic motion is key to reducing the overall vibration of the machine. Full article
(This article belongs to the Section Agricultural Technology)
Show Figures

Figure 1

Figure 1
<p>The 3D model structure of the pickup baler.</p>
Full article ">Figure 2
<p>The operation principle and harvesting process of the machine.</p>
Full article ">Figure 3
<p>Straw harvesting process of the pickup baler. (<b>a</b>) Manual bale stacking; (<b>b</b>) bale unloading at the field edge.</p>
Full article ">Figure 4
<p>Schematic of the pickup header device. (<b>a</b>) Actual image of pickup header; (<b>b</b>) force diagram of the header and conveyor trough positions.</p>
Full article ">Figure 5
<p>Mechanical analysis diagram of crop roller. (<b>a</b>) The front view of crop roller; (<b>b</b>) force model of crop roller.</p>
Full article ">Figure 6
<p>Interaction forces between connected kinematic pairs. (<b>a</b>) Force schematic diagram of the crank; (<b>b</b>) force schematic diagram of the connecting rod; (<b>c</b>) force schematic diagram of the piston.</p>
Full article ">Figure 7
<p>1A312E type accelerometer.</p>
Full article ">Figure 8
<p>DH5902N type dynamic signal acquisition system.</p>
Full article ">Figure 9
<p>Vibration test testing scenarios. (<b>a</b>) Engine position measurement point arrangement; (<b>b</b>) indoor vibration testing.</p>
Full article ">Figure 10
<p>The indoor measurement point arrangement.</p>
Full article ">Figure 11
<p>The outdoor measurement point arrangement.</p>
Full article ">Figure 12
<p>The heat map of vibration intensity in time domain under different operating conditions of the self-propelled straw pickup baler. (<b>a</b>) RMS value of operating condition 1; (<b>b</b>) RMS value of operating condition 2; (<b>c</b>) peak-to-peak value of operating condition 1; (<b>d</b>) peak-to-peak value of operating condition 2.</p>
Full article ">Figure 13
<p>Comparison of sensor layout and RMS envelope spectrum signal. (<b>a</b>) Sensor layout diagram for measurement point 7; (<b>b</b>) measurement point 7 (condition 1); (<b>c</b>) measurement point 7 (condition 2).</p>
Full article ">Figure 14
<p>Time domain response of acceleration signals. (<b>a</b>) Measurement point 2; (<b>b</b>) measurement point 8.</p>
Full article ">Figure 15
<p>Comparison of vibration intensity in time domain under different conditions for the pickup baler in outdoor testing. (<b>a</b>) Positive and negative amplitude of vibration acceleration under test condition 3; (<b>b</b>) positive and negative amplitude of vibration acceleration under test condition 4.</p>
Full article ">Figure 16
<p>Idle throttle engine frequency spectrum (X direction). (<b>a</b>) First ten peak frequencies; (<b>b</b>) low-frequency band enlarged display.</p>
Full article ">Figure 17
<p>The frequencies of each measurement 1~6 point under condition 1.</p>
Full article ">Figure 18
<p>The frequencies of each measurement 7~12 point under condition 1.</p>
Full article ">Figure 19
<p>Low-frequency component frequency spectrum. (<b>a</b>) Measurement point 1-X; (<b>b</b>) measurement point 2-Y; (<b>c</b>) measurement point 6-X; (<b>d</b>) measurement point 7-Z; (<b>e</b>) measurement point 8-Z; (<b>f</b>) measurement point 12-Z.</p>
Full article ">Figure 20
<p>Comparison diagram of signal filtering before and after filtering at measuring point 8 on the left rear of the chassis frame.</p>
Full article ">Figure 21
<p>Amplitude spectrum waterfall plots of each measurement point. (<b>a</b>) Time–frequency map of 1-X point; (<b>b</b>) time–frequency map of 2-Y point; (<b>c</b>) time–frequency map of 3-X point; (<b>d</b>) time–frequency map of 4-Z point; (<b>e</b>) time–frequency map of 5-Z point; (<b>f</b>) time–frequency map of 6-X point.</p>
Full article ">Figure 21 Cont.
<p>Amplitude spectrum waterfall plots of each measurement point. (<b>a</b>) Time–frequency map of 1-X point; (<b>b</b>) time–frequency map of 2-Y point; (<b>c</b>) time–frequency map of 3-X point; (<b>d</b>) time–frequency map of 4-Z point; (<b>e</b>) time–frequency map of 5-Z point; (<b>f</b>) time–frequency map of 6-X point.</p>
Full article ">
23 pages, 19563 KiB  
Review
A Review: Phase Measurement Techniques Based on Metasurfaces
by Zhicheng Zhao, Yueqiang Hu and Shanyong Chen
Photonics 2024, 11(11), 996; https://doi.org/10.3390/photonics11110996 - 22 Oct 2024
Viewed by 1262
Abstract
Phase carries crucial information about the light propagation process, and the visualization and quantitative measurement of phase have important applications, ranging from ultra-precision metrology to biomedical imaging. Traditional phase measurement techniques typically require large and complex optical systems, limiting their applicability in various [...] Read more.
Phase carries crucial information about the light propagation process, and the visualization and quantitative measurement of phase have important applications, ranging from ultra-precision metrology to biomedical imaging. Traditional phase measurement techniques typically require large and complex optical systems, limiting their applicability in various scenarios. Optical metasurfaces, as flat optical elements, offer a novel approach to phase measurement by manipulating light at the nanoscale through light-matter interactions. Metasurfaces are advantageous due to their lightweight, multifunctional, and easy-to-integrate nature, providing new possibilities for simplifying traditional phase measurement methods. This review categorizes phase measurement techniques into quantitative and non-quantitative methods and reviews the advancements in metasurface-based phase measurement technologies. Detailed discussions are provided on several methods, including vortex phase contrast, holographic interferometry, shearing interferometry, the Transport of Intensity Equation (TIE), and wavefront sensing. The advantages and limitations of metasurfaces in phase measurement are highlighted, and future research directions are explored. Full article
(This article belongs to the Special Issue Challenges and Future Directions in Adaptive Optics Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the concept for spin-dependent function control. (<b>b</b>) Schematic of the designed all-dielectric metasurface spatial filter. (<b>c</b>) Traditional bright field and spiral phase contrast images of the undyed onion epidermal cells captured with LCP and RCP incident light at the wavelength of 480, 530, 580, and 630 nm [<a href="#B49-photonics-11-00996" class="html-bibr">49</a>]. (<b>d</b>) Schematic illustration of the spiral metalens with a simplified optical system. (<b>e</b>) Unit cell structure description with tilted and top views. (<b>f</b>) Bright-field images of erythrocytes with ×50 objective lens and edge-enhanced images with the spiral metalens at 497, 532, 580, and 633 nm wavelengths [<a href="#B50-photonics-11-00996" class="html-bibr">50</a>]. (<b>g</b>) Sketch of the experimental setup for simultaneous spiral phase contrast and bright-field imaging. (<b>h</b>) Schematic diagram of the designed dielectric metasurface for synchronously spiral phase contrast and bright-field imaging. (<b>i</b>) Synchronously captured spiral phase contrast and bright field images of “META” and unstained limewood stem cells in the same field of view [<a href="#B51-photonics-11-00996" class="html-bibr">51</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic of the Nanophotonics Enhanced Coverslip (NEC) phase image system at 637 nm. (<b>b</b>) Schematic of the NEC. (<b>c</b>–<b>e</b>) Phase imaging of HeLa cells with NEC, conventional DIC and fluorescence [<a href="#B52-photonics-11-00996" class="html-bibr">52</a>]. (<b>f</b>) Schematic of phase imaging using spin-orbit coupling enabled by plasmonic metasurface. (<b>g</b>–<b>h</b>) Phase imaging of the eggcrate pattern with metasurface and without metasurface [<a href="#B53-photonics-11-00996" class="html-bibr">53</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) The schematic for high-resolution, widefield measurement of phase alterations introduced by plasmonic metasurfaces. The metasurface acts as a geometric phase grating (GPG). (<b>b</b>) Amplitude and phase image of a vortex LG beam metasurface [<a href="#B56-photonics-11-00996" class="html-bibr">56</a>]. (<b>c</b>) Schematic of the metasurface, which is composed of rectangular TiO<sub>2</sub> nanopillars on a fused silica substrate. (<b>d</b>) Schematic of common path digital holographic system for quantitative phase imaging with a singlelayer metasurface. (<b>e</b>) Experimental demonstration of digital holography on test object: object, image plane hologram, phase map, and the height along white line [<a href="#B57-photonics-11-00996" class="html-bibr">57</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic of the QPGM employing two metasurface layers. (<b>b</b>) Schematics of a uniform array of rectangular nanoposts (top) and a single unit cell (bottom). (<b>c</b>) Optical images of the fabricated metasurfaces. (<b>d</b>) Thicknesses of seven different phase targets calculated by the QPGM, and those measured by AFM. (<b>e</b>) Schematic of a sea urchin cell and its corresponding phase gradient images. Scale bars, 40 μm [<a href="#B58-photonics-11-00996" class="html-bibr">58</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Experiment setup of the proposed FOSSM. Obj, object. P, polarizer. L, lens.MS, metasurface. A, analyzer. (<b>b</b>) The concept of retardance imaging of the object with a laterally (along the x direction) and a longitudinally (along the z direction) displaced metasurface. (<b>c</b>) Quantitative phase imaging of NIH3T3 cells with a laterally displaced metasurface [<a href="#B59-photonics-11-00996" class="html-bibr">59</a>]. (<b>d</b>) Schematic of single-shot quantitative amplitude and phase imaging based on a pair of dielectric geometric phase metasurfaces. (<b>e</b>) Designed geometric phases of two metasurfaces. (<b>f</b>) Amplitude and phase of the object reconstructed by using a series of retardance images. (<b>g</b>) Recovered amplitude and phase of SKNO-1 cells [<a href="#B60-photonics-11-00996" class="html-bibr">60</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) The experimental setup optical vector differential operations based on computing metasurfaces. Path 1 in the Mach−Zehnder interferometer performs the differential operation in the x direction, and path 2 does the differential operation in the y direction. (<b>b</b>) Imaging results of fishtail cross-cut cells with broadband vectorial DIC microscopy. Bright-field images and DIC microscopy images for different wavelengths [<a href="#B61-photonics-11-00996" class="html-bibr">61</a>]. (<b>c</b>) Schematic of the metalens-assisted single-shot complex amplitude imaging system. Captured x and y shearing interference patterns with the polarization channel along 0°, 45°, 90°, and 135°, respectively. (<b>d</b>) Calculated phase gradients along the x and y direction, respectively. (<b>e</b>) Surface morphology of UV adhesive measured by the metalens-assisted system and a commercial white light interferometer (WLI) [<a href="#B62-photonics-11-00996" class="html-bibr">62</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic of the 2D edge detection and metasurface. The light incidents onto the “EDGE” shaped object, and then passes through the metasurface at the Fourier plane, and finally, its edge information is obtained at the image plane. (<b>b</b>) Edge detection of the human umbilical vein endothelial cell (first row) and bronchial epithelial cell (second row). The imaging methods are bright field, phase contrast, dark field and bright field and edge detection from left to right successively [<a href="#B63-photonics-11-00996" class="html-bibr">63</a>]. (<b>c</b>) Principle of metasurface-assisted i-DIC microscopy and Si meta-atom. (<b>d</b>) Imaging results with a-DIC and i-DIC microscopy [<a href="#B64-photonics-11-00996" class="html-bibr">64</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic diagram of the metasurface (MS)-based quantitative phase imaging setup (MS-TIE). L1 and L2 form a 4f optical setup. The metasurface is placed at the Fourier plane and acts as a polarization-dependent optical filter. (<b>b</b>) Unit cell of the metasurface consisting of amorphous silicon nanopillars on a fused silica substrate. (<b>c</b>) Contrast phase imaging error between metasurfaces and traditional TIE method [<a href="#B65-photonics-11-00996" class="html-bibr">65</a>]. (<b>d</b>) Schematic diagram of triple Transport of Intensity Equation phase retrieval based on anisotropic metasurface. One image is in focus, two images are defocus, and the defocus distance is fixed and conjugate. (<b>e</b>) Schematic of meta-atom and scanning electron microscope (SEM) image of metasurface. (<b>f</b>) The experimental phase-only object results. Target phase map, the single-shot captured triple images via metasurface, and the reconstructed phase image based on TTIE algorithm [<a href="#B66-photonics-11-00996" class="html-bibr">66</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematics of the dispersive metalens-based QPI. (<b>b</b>) Photograph of the meta-microscope (length: 36 mm, width: 36 mm, and height: 14 mm). (<b>c</b>) Measured intensity distributions of the longitudinal light-field cross-sections at targeted wavelengths. (<b>d</b>) Reconstructed in-focus phase profiles of the 4T1 cells from the image stack obtained by the meta-microscope. Scale bar is 20 μm [<a href="#B67-photonics-11-00996" class="html-bibr">67</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) Measurement protocol based on asymmetric metasurface photodetectors, where the sensor array is partitioned into blocks of four adjacent pixels coated with the asymmetric metasurface oriented along four orthogonal directions. (<b>b</b>) Reconstructed phase distribution of the MCF-10A cell [<a href="#B68-photonics-11-00996" class="html-bibr">68</a>]. (<b>c</b>) Schematic of phase imaging system using non-local metasurfce (NLM). (<b>d</b>) Phase imaging results with Zernike’s method and NLM [<a href="#B69-photonics-11-00996" class="html-bibr">69</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Design principle of meta Shack–Hartmann wavefront sensor. (<b>b</b>) Phase imaging results: object and reconstructed phase [<a href="#B71-photonics-11-00996" class="html-bibr">71</a>]. (<b>c</b>) The wavefront sensor consists of a CCD, an MA, and two linear polarizers and can operate at both 950 nm and 1030 nm. (<b>d</b>) Experimental demonstration of the spot centroid shift in the x-y plane and the corresponding reconstructed wavefront at 950 nm [<a href="#B72-photonics-11-00996" class="html-bibr">72</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) Schematics of experimental setup of phase and amplitude reconstruction by weak measurement. (<b>b</b>) Phase imaging results: object and reconstructed phase [<a href="#B76-photonics-11-00996" class="html-bibr">76</a>]. (<b>c</b>) Schematic illustration of computational complex field retrieval using a designed metasurface diffuser (MD). (<b>d</b>) Phase imaging results: object and reconstructed phase [<a href="#B77-photonics-11-00996" class="html-bibr">77</a>].</p>
Full article ">
17 pages, 3863 KiB  
Article
One-Dimensional Numerical Cascade Model of Runoff and Soil Loss on Convergent and Divergent Plane Soil Surfaces: Laboratory Assessment and Numerical Simulations
by Babar Mujtaba, João L. M. P. de Lima and M. Isabel P. de Lima
Water 2024, 16(20), 2955; https://doi.org/10.3390/w16202955 - 17 Oct 2024
Viewed by 618
Abstract
A one-dimensional numerical overland flow model based on the cascade plane theory was developed to estimate rainfall-induced runoff and soil erosion on converging and diverging plane surfaces. The model includes three components: (i) soil infiltration using Horton’s infiltration equation, (ii) overland flow using [...] Read more.
A one-dimensional numerical overland flow model based on the cascade plane theory was developed to estimate rainfall-induced runoff and soil erosion on converging and diverging plane surfaces. The model includes three components: (i) soil infiltration using Horton’s infiltration equation, (ii) overland flow using the kinematic wave approximation of the one-dimensional Saint-Venant shallow water equations for a cascade of planes, and (iii) soil erosion based on the sediment transport continuity equation. The model’s performance was evaluated by comparing numerical results with laboratory data from experiments using a rainfall simulator and a soil flume. Four independent experiments were conducted on converging and diverging surfaces under varying slope and rainfall conditions. Overall, the numerically simulated hydrographs and sediment graphs closely matched the laboratory results, showing the efficiency of the model for the tested controlled laboratory conditions. The model was then used to numerically explore the impact of different plane soil surface geometries on runoff and soil loss. Seven geometries were studied: one rectangular, three diverging, and three converging. A constant soil surface area, the rainfall intensity, and the slope gradient were maintained in all simulations. Results showed that increasing convergence angles led to a higher peak and total soil loss, while decreasing divergence angles reduced them. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch and notation used for the cascade of <span class="html-italic">n</span>-planes representing converging (sloping to the left) or diverging (sloping to the right) cascades.</p>
Full article ">Figure 2
<p>(<b>a</b>) Laboratory setup used in the experiments, consisting of a rainfall simulator and a soil flume (top and bottom left); the flume has outlets on both ends for the converging plane (bottom middle) and diverging plane surface (bottom right) experiments. (<b>b</b>) Dashed lines represent contour lines relative to an arbitrary datum (ground level) for a 20% flume slope, while solid lines indicate the border wall of the flume’s geometry.</p>
Full article ">Figure 3
<p>On the <b>right</b>, a schematic sketch illustrates the approximate representation of the soil flume surface planar geometry using converging and diverging cascade planes, each with an equal length <math display="inline"><semantics> <mrow> <mi>L</mi> </mrow> </semantics></math> of 0.5 m, measured along the direction of the plane slope. The width <math display="inline"><semantics> <mrow> <mi>W</mi> </mrow> </semantics></math> of each plane is also shown, which is measured in the perpendicular direction. On the <b>left</b>, the variation in mean rainfall intensity across the planes of the converging and diverging cascades is shown.</p>
Full article ">Figure 4
<p>Seven plane soil surfaces: one rectangular, three converging with different convergence angles θ, and three diverging with different divergence angles θ. The schematic sketch also illustrates the approximate representation of the soil surface planar geometry using converging and diverging cascade planes of equal length <span class="html-italic">L</span> (1 m). The width <span class="html-italic">W</span> of each plane is also shown. The dimensions of the rectangular plane soil surface (4 m × 2 m) are provided as well.</p>
Full article ">Figure 5
<p>Observed and numerically simulated hydrographs for converging and diverging surfaces with different combinations of rainfall intensity (<span class="html-italic">I</span>) and slope (<span class="html-italic">S</span>). The Nash–Sutcliffe (<math display="inline"><semantics> <mrow> <mi>N</mi> <mi>S</mi> </mrow> </semantics></math>) coefficients are also shown.</p>
Full article ">Figure 6
<p>Observed and numerically simulated sediment graphs for converging and diverging surfaces with different combinations of rainfall intensity (<span class="html-italic">I</span>) and slope (<span class="html-italic">S</span>). Note that the Y-axis scale is not the same on the top-left figure. The Nash–Sutcliffe (<math display="inline"><semantics> <mrow> <mi>N</mi> <mi>S</mi> </mrow> </semantics></math>) coefficients are also shown.</p>
Full article ">Figure 7
<p>Simulated hydrographs for plane soil surfaces with different angles (θ) of convergence (<b>left</b>) and divergence (<b>right</b>). Section A represents a rectangular surface where θ = 0°.</p>
Full article ">Figure 8
<p>Runoff peaks for plane soil surfaces with different angles (θ) of convergence (<b>left</b>) and divergence (<b>right</b>). A rectangular surface is represented by θ = 0°. Trend lines fitted to the data are included for reference.</p>
Full article ">Figure 9
<p>Simulated sediment graphs for plane soil surfaces with different angles (θ) of convergence (<b>left</b>) and divergence (<b>right</b>). Section A represents a rectangular surface where θ = 0°. Note that the Y-axis scales are not the same.</p>
Full article ">Figure 10
<p>Peak soil loss for plane soil surfaces with different angles (θ) of convergence (<b>left</b>) and divergence (<b>right</b>). A rectangular surface is represented by θ = 0°. Trend lines fitted to the data are included for reference. Note that the Y-axis scales differ.</p>
Full article ">
20 pages, 4725 KiB  
Article
A Light-Powered Self-Circling Slider on an Elliptical Track with a Liquid Crystal Elastomer Fiber
by Lu Wei, Yanan Chen, Junjie Hu, Xueao Hu, Jiale Wang and Kai Li
Polymers 2024, 16(16), 2375; https://doi.org/10.3390/polym16162375 - 22 Aug 2024
Viewed by 759
Abstract
In this paper, we propose an innovative light-powered LCE-slider system that enables continuous self-circling on an elliptical track and is comprised of a light-powered LCE string, slider, and rigid elliptical track. By formulating and solving dimensionless dynamic equations, we explain static and self-circling [...] Read more.
In this paper, we propose an innovative light-powered LCE-slider system that enables continuous self-circling on an elliptical track and is comprised of a light-powered LCE string, slider, and rigid elliptical track. By formulating and solving dimensionless dynamic equations, we explain static and self-circling states, emphasizing self-circling dynamics and energy balance. Quantitative analysis reveals that the self-circling frequency of LCE-slider systems is independent of the initial tangential velocity but sensitive to light intensity, contraction coefficients, elastic coefficients, the elliptical axis ratio, and damping coefficients. Notably, elliptical motion outperforms circular motion in angular velocity and frequency, indicating greater efficiency. Reliable self-circling under constant light suggests applications in periodic motion fields, especially celestial mechanics. Additionally, the system’s remarkable adaptability to a wide range of curved trajectories exemplifies its flexibility and versatility, while its energy absorption and conversion capabilities position it as a highly potential candidate for applications in robotics, construction, and transportation. Full article
(This article belongs to the Special Issue Polymer Materials for Sensors and Actuators)
Show Figures

Figure 1

Figure 1
<p>Diagram of a light-powered self-circling device on a plane including the LCE fiber, sliding element, and fixed elliptical track: (<b>a</b>) Initial state; (<b>b</b>) Current state; and (<b>c</b>) Force analysis. Under constant light exposure, the slider equipped with the LCE fiber achieves autonomous, sustained, and periodic movement along the elliptical track.</p>
Full article ">Figure 2
<p>Two motion regimes of the system under continuous light exposure: static state and self-circling state. (<b>a</b>,<b>b</b>) Time-dependent diagram of angular displacement at <math display="inline"><semantics> <mrow> <mover> <mi>I</mi> <mo>-</mo> </mover> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>c</b>) Angular velocity dependence on angular displacement at <math display="inline"><semantics> <mrow> <mover> <mi>I</mi> <mo>-</mo> </mover> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>; (<b>d</b>,<b>e</b>) Time-dependent diagram of angular displacement at <math display="inline"><semantics> <mrow> <mover> <mi>I</mi> <mo>-</mo> </mover> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>; (<b>f</b>) Angular velocity dependence on angular displacement <math display="inline"><semantics> <mrow> <mover> <mi>I</mi> <mo>-</mo> </mover> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Self-circling mechanism of the system. (<b>a</b>) Time-dependent curve of angular displacement; (<b>b</b>) Time-dependent curve of the number fraction of <span class="html-italic">cis</span>-<span class="html-italic">isomers</span> in the LCE fiber; (<b>c</b>) Time-dependent curve of tangential tension in the LCE fiber; (<b>d</b>) Time-dependent curve of damping force; (<b>e</b>) Variation of tangential tension in the LCE fiber with respect to the angular displacement; (<b>f</b>) Variation of damping force with respect to angular displacement. The yellow squares in the figure denote illuminated region.</p>
Full article ">Figure 4
<p>The effect of light intensity on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover> <mi>I</mi> <mo>-</mo> </mover> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>,<math display="inline"><semantics> <mrow> <mo> </mo> <mn>0.45</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.6</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with light intensities.</p>
Full article ">Figure 5
<p>The effect of the contraction coefficient on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mi>C</mi> <mo>=</mo> <mn>0.3</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.35</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>0.4</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with the contraction coefficient.</p>
Full article ">Figure 6
<p>The effect of the elastic coefficient on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>K</mi> <mo>¯</mo> </mover> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>1.0</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>1.3</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with the elastic coefficient.</p>
Full article ">Figure 7
<p>The effect of the initial tangential velocity on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>v</mi> <mn>0</mn> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> <mo>=</mo> <mn>0.9</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>1.2</mn> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>1.5</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with initial tangential velocity.</p>
Full article ">Figure 8
<p>The effect of damping coefficients on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>β</mi> <mn>1</mn> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.007</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>0.008</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with the first-order damping coefficient <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>1</mn> </msub> </mrow> </semantics></math>. (<b>c</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <msub> <mi>β</mi> <mn>2</mn> </msub> </mrow> <mo stretchy="true">¯</mo> </mover> <mo>=</mo> <mn>0.0003</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>0.0006</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>0.0009</mn> </mrow> </semantics></math>. (<b>d</b>) Self-circling frequency variations with the second-order damping coefficient <math display="inline"><semantics> <mrow> <msub> <mi>β</mi> <mn>2</mn> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>The effect of the elliptical semi-major axis on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>a</mi> <mo>¯</mo> </mover> <mo>=</mo> <mn>1.92</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>2.00</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>2.08</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with elliptical semi-major axis.</p>
Full article ">Figure 10
<p>The effect of the elliptical semi-minor axis on the self-circling frequency. (<b>a</b>) Depictions of limit cycles at <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>b</mi> <mo>¯</mo> </mover> <mo>=</mo> <mn>1.84</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>1.90</mn> <mo>,</mo> <mo> </mo> </mrow> </semantics></math>and <math display="inline"><semantics> <mrow> <mn>1.96</mn> </mrow> </semantics></math>. (<b>b</b>) Self-circling frequency variations with elliptical semi-minor axis.</p>
Full article ">
15 pages, 879 KiB  
Article
Solar Energetic Particles Propagation under 3D Corotating Interaction Regions with Different Characteristic Parameters
by Yuji Zhu and Fang Shen
Universe 2024, 10(8), 315; https://doi.org/10.3390/universe10080315 - 2 Aug 2024
Cited by 1 | Viewed by 873
Abstract
Solar energetic particles (SEPs) are bursts of high-energy particles that originate from the Sun and can last for hours or even days. The aim of this study is to understand how the characteristics of energetic particles ware affected by the characteristic parameters of [...] Read more.
Solar energetic particles (SEPs) are bursts of high-energy particles that originate from the Sun and can last for hours or even days. The aim of this study is to understand how the characteristics of energetic particles ware affected by the characteristic parameters of corotating interaction regions (CIRs). In particular, the particle intensity distribution with time and space in CIRs with different characteristics were studied. The propagation and acceleration of particles were described by the focused transport equation (FTE). We used a three-dimensional magnetohydrodynamic (MHD) model to simulate the background solar wind with CIRs. By changing the inner boundary conditions, we constructed CIRs with different solar wind speeds, angles between the polar axis and rotation axis, and the azimuthal widths of the fast streams. Particles were impulsively injected at the inner boundary of the MHD model. We then studied the particle propagation and compression acceleration in different background solar wind. The results showed that the CIR widths are related to the solar wind speed, tilt angles, and the azimuthal widths of the fast stream. The acceleration of particles in the reverse and forward compression regions are mainly influenced by the solar wind speed difference and the slow solar wind speed, respectively. Particles with lower energy (sub-MeV) are more sensitive to the solar wind speed difference and the tilt angle. The particle intensity variation with time and the radial distance is mainly influenced by the solar wind speed. The longitudinal distribution of particle intensity is affected by the solar wind speed, tilt angles, and the azimuthal widths of the fast stream. Full article
(This article belongs to the Section Space Science)
Show Figures

Figure 1

Figure 1
<p>The radial solar wind speed at the inner boundary of the cases in <b>Set A</b>–<b>D</b>. The cases in <b>Set A</b> have the same <math display="inline"><semantics> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msubsup> </semantics></math>. The cases in <b>Set B</b> have the same <math display="inline"><semantics> <mrow> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> <mo>−</mo> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msubsup> </mrow> </semantics></math>. The cases in <b>Set C</b> have the same <math display="inline"><semantics> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> </semantics></math>, but different <math display="inline"><semantics> <mi>α</mi> </semantics></math>. The cases in <b>Set D</b> have the same <math display="inline"><semantics> <mi>α</mi> </semantics></math>, <math display="inline"><semantics> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msubsup> </semantics></math>, and <math display="inline"><semantics> <msubsup> <mi>V</mi> <mrow> <mi>b</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msubsup> </semantics></math>, but different <math display="inline"><semantics> <mrow> <mi>δ</mi> <msub> <mi>ϕ</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>. From panel (<b>D1</b>–<b>D3</b>), <math display="inline"><semantics> <mrow> <mi>δ</mi> <msub> <mi>ϕ</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>r</mi> <mi>e</mi> <mi>a</mi> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math> is <math display="inline"><semantics> <mrow> <mn>136</mn> <mo>°</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mn>45</mn> <mo>°</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mn>18</mn> <mo>°</mo> </mrow> </semantics></math>, respectively.</p>
Full article ">Figure 2
<p>The CIR width of the selected cases. The blue, red and yellow asterisks indicate the CIR radial extent, width projected into the solar equatorial plane and CIR width considering 3D geometry, respectively. Panels (<b>A</b>–<b>D</b>) are for cases in <b>Set A</b>–<b>D</b>, respectively.</p>
Full article ">Figure 3
<p>The ratio of the maximum energy to the original energy (<math display="inline"><semantics> <msub> <mi>E</mi> <mn>0</mn> </msub> </semantics></math> = 5 MeV) at 1 AU after 60 h versus the CIR parameters. The red asterisks and blue circles are for the reverse and the forward compression regions, respectively. Panels (<b>A</b>–<b>D</b>) are for cases in <b>Set A</b>–<b>D</b>, respectively.</p>
Full article ">Figure 4
<p>The divergence of the background solar wind velocity <math display="inline"><semantics> <mrow> <mo>∇</mo> <mo>·</mo> <mi>V</mi> </mrow> </semantics></math> of Case 1 (panel (<b>a</b>)), Case 3 (panel (<b>b</b>)), and Case 5 (panel (<b>c</b>)) in Set B.</p>
Full article ">Figure 5
<p>The maximum ratio of the maximum energy in the reversed compression region to that in the approximated Parker region at 1 AU in the first 60 h versus the CIR parameters. The blue asterisks, red circles, and yellow squares are for protons with original energy of 0.5 Mev, 5 MeV, and 20 MeV, respectively. Panels (<b>A</b>–<b>D</b>) are for cases in <b>Set A</b>–<b>D</b>, respectively.</p>
Full article ">Figure 6
<p>The peak intensity of particles versus the CIR radial extent <math display="inline"><semantics> <msub> <mi>D</mi> <mi>c</mi> </msub> </semantics></math> at 1 AU. The blue asterisks, red circles, yellow squares, and purple crosses indicate cases in <b>Set A</b>, <b>Set B</b>, <b>Set C</b>, and <b>Set D</b>, respectively.</p>
Full article ">Figure 7
<p>The variations of particle intensity (6.7–11.4 MeV) with time, longitude, and radial distance for different cases. The rows from top to bottom display the cases in <b>Set A</b>–<b>D</b>, respectively. The columns from left to right represent the variations of particle intensity with time, longitude, and radial distance, respectively. Panels <b>A3(a)</b>–<b>D3(a)</b> and panels <b>A3(b)</b>–<b>D3(b)</b> show the distribution after 30 h and 120 h since the injection, respectively.</p>
Full article ">
18 pages, 5340 KiB  
Article
Measurement and Calculation of Sediment Transport on an Ephemeral Stream
by Loukas Avgeris, Konstantinos Kaffas and Vlassios Hrissanthou
Hydrology 2024, 11(7), 96; https://doi.org/10.3390/hydrology11070096 - 30 Jun 2024
Viewed by 974
Abstract
Sediment transport remains a significant challenge for researchers due to the intricate nature of the physical processes involved and the diverse characteristics of watercourses worldwide. A type of watercourse that is of particular interest for study is the ephemeral streams, found primarily in [...] Read more.
Sediment transport remains a significant challenge for researchers due to the intricate nature of the physical processes involved and the diverse characteristics of watercourses worldwide. A type of watercourse that is of particular interest for study is the ephemeral streams, found primarily in semiarid and arid regions. Due to their unique nature, a new measurement algorithm was created and a modified bed load sampler was built. Measurement of the bed load transport rate and calculation of the water discharge were conducted in an ephemeral stream in Northeastern Greece, where the mean calculated streamflow rate ranged from 0.019 to 0.314 m3/s, and the measured sediment load transport rates per unit width varied from 0.00001 to 0.00213 kg/m/s. The sediment concentration was determined through various methods, including nonlinear regression equations and formulas developed by Yang, with the coefficients of these formulas calibrated accordingly. The results demonstrated that the equations derived from Yang’s multiple regression analysis offered a superior fit compared to the original equations. As a result, two modified versions of Yang’s stream sediment transport formulas were developed and are presented to the readership. To assess the accuracy of the modified formulas, a comparison was conducted between the calculated total sediment concentrations and the measured total sediment concentrations based on various statistical criteria. The analysis shows that none of Yang’s original formulas fit the available data well, but after optimization, both modified formulas can be applied to the specific ephemeral stream. The results indicate also that the formulas derived from the nonlinear regression can be successfully used for the determination of the total sediment concentration in the ephemeral stream and have a better fit compared to Yang’s formulas. The correlation from the nonlinear regression equations suggests that total sediment transport is primarily influenced by water discharge and rainfall intensity, with the latter showing a high correlation coefficient of 0.998. Full article
(This article belongs to the Special Issue Advances in Catchments Hydrology and Sediment Dynamics)
Show Figures

Figure 1

Figure 1
<p>Study area: (<b>a</b>) placement on a map of Greece; (<b>b</b>) placement on Kosynthos basin; (<b>c</b>) ephemeral stream basin.</p>
Full article ">Figure 2
<p>Parts of the trap: (<b>a</b>) schematic diagram; (<b>b</b>) metal frame, ground plate with inclined front edge, metal stakes, and nylon netting with webbing straps.</p>
Full article ">Figure 3
<p>The trap installed: (<b>a</b>) before the event; (<b>b</b>) after the event.</p>
Full article ">Figure 4
<p>Flowchart of the measurement’s algorithm.</p>
Full article ">Figure 5
<p>Sediment size distribution curve.</p>
Full article ">Figure 6
<p>Streamflow hydrograph and precipitation hyetograph (January 2017–July 2018).</p>
Full article ">Figure 7
<p>Discrepancy ratio plot between measured and calculated values of total sediment concentration in the ephemeral stream by means of the original and the calibrated Yang formula (1973).</p>
Full article ">Figure 8
<p>Discrepancy ratio plot between measured and calculated values of total sediment concentration in the ephemeral stream by means of the original and the calibrated Yang formula (1979).</p>
Full article ">Figure 9
<p>(<b>a</b>) Discrepancy ratio plot between measured and calculated values of total sediment concentration in the ephemeral stream (<b>a</b>) based on the water discharge and rainfall intensity and (<b>b</b>) based on the combination of the water discharge and rainfall intensity with median particle size.</p>
Full article ">
22 pages, 12698 KiB  
Article
Non-Equilibrium Scour Evolution around an Emerged Structure Exposed to a Transient Wave
by Deniz Velioglu Sogut, Erdinc Sogut, Ali Farhadzadeh and Tian-Jian Hsu
J. Mar. Sci. Eng. 2024, 12(6), 946; https://doi.org/10.3390/jmse12060946 - 5 Jun 2024
Cited by 1 | Viewed by 909
Abstract
The present study evaluates the performance of two numerical approaches in estimating non-equilibrium scour patterns around a non-slender square structure subjected to a transient wave, by comparing numerical findings with experimental data. This study also investigates the impact of the structure’s positioning on [...] Read more.
The present study evaluates the performance of two numerical approaches in estimating non-equilibrium scour patterns around a non-slender square structure subjected to a transient wave, by comparing numerical findings with experimental data. This study also investigates the impact of the structure’s positioning on bed evolution, analyzing configurations where the structure is either attached to the sidewall or positioned at the centerline of the wave flume. The first numerical method treats sediment particles as a distinct continuum phase, directly solving the continuity and momentum equations for both sediment and fluid phases. The second method estimates sediment transport using the quadratic law of bottom shear stress, yielding robust predictions of bed evolution through meticulous calibration and validation. The findings reveal that both methods underestimate vortex-induced near-bed vertical velocities. Deposits formed along vortex trajectories are overestimated by the first method, while the second method satisfactorily predicts the bed evolution beneath these paths. Scour holes caused by wave impingement tend to backfill as the flow intensity diminishes. The second method cannot sufficiently capture this backfilling, whereas the first method adequately reflects the phenomenon. Overall, this study highlights significant variations in the predictive capabilities of both methods in regard to the evolution of non-equilibrium scour at low Keulegan–Carpenter numbers. Full article
(This article belongs to the Special Issue Coastal Disaster Assessment and Response)
Show Figures

Figure 1

Figure 1
<p>Experimental setup and plan view of WGs and ADVs for (<b>a</b>) side and (<b>b</b>) center layouts. Yellow circle and gray triangle represent WG and ADV, respectively.</p>
Full article ">Figure 2
<p>Grid sensitivity analyses based on the horizontal velocity component, <math display="inline"><semantics> <mrow> <mi>u</mi> </mrow> </semantics></math>, captured by ADV1 and ADV3 for side layout: (<b>a</b>) SedWaveFoam; (<b>b</b>) FLOW-3D.</p>
Full article ">Figure 3
<p>Final mesh of NWF for both models. (<b>a</b>) side; (<b>b</b>) center layouts on <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> <mi>y</mi> </mrow> </semantics></math> plane; (<b>c</b>) both layouts on <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> <mi>z</mi> </mrow> </semantics></math> plane. All units are in mm. Not to scale.</p>
Full article ">Figure 4
<p>Comparison of measured (gray circles) and computed free surface elevations (<math display="inline"><semantics> <mrow> <mi>η</mi> </mrow> </semantics></math>) and velocity components: (<b>a</b>–<b>c</b>) side; (<b>b</b>–<b>d</b>) center layout. Red and black solid lines represent SedWaveFoam and FLOW-3D results, respectively.</p>
Full article ">Figure 5
<p>Plan view of sandy bed at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s. Upper panels: side layout. Lower panels: center layout. The figure compares the measured and predicted bed elevations using three different <math display="inline"><semantics> <mrow> <mi>α</mi> </mrow> </semantics></math> values.</p>
Full article ">Figure 6
<p>Plan view of sandy bed at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s. (<b>a</b>) Measurement; FLOW-3D results obtained using equations of (<b>b</b>) Nielsen; (<b>c</b>) van Rjin; (<b>d</b>) Meyer-Peter–Müller.</p>
Full article ">Figure 7
<p>Plan view of sandy bed at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s. (<b>a</b>) Measurement; FLOW-3D results for different turbulence schemes: (<b>b</b>) LES; (<b>c</b>) RNG; (<b>d</b>) k − ε; (<b>e</b>) k − ω.</p>
Full article ">Figure 8
<p>Spatial distributions of <math display="inline"><semantics> <mrow> <mi>U</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>w</mi> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 7 s and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 8 s, extracted from (<b>a</b>,<b>c</b>) <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math> m above bed; (<b>b</b>,<b>d</b>) bed surface. (<b>a</b>,<b>b</b>) and (<b>c</b>,<b>d</b>) are SedWaveFoam and FLOW-3D results, respectively. Upper two panels: side layout. Lower two panels: center layout.</p>
Full article ">Figure 9
<p>Spatial distributions of <math display="inline"><semantics> <mrow> <mi>U</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>z</mi> </mrow> </msub> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 7 s and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 8 s, extracted from (<b>a</b>,<b>c</b>) <math display="inline"><semantics> <mrow> <mi>h</mi> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math> m above bed; (<b>b</b>,<b>d</b>) bed surface. (<b>a</b>,<b>b</b>) and (<b>c</b>,<b>d</b>) are SedWaveFoam and FLOW-3D results, respectively. Upper two panels: side layout. Lower two panels: center layout.</p>
Full article ">Figure 10
<p>Bed elevation and suspended sediment concentration (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ϕ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msup> <mo>)</mo> </mrow> </semantics></math> in the vicinity of the structure at various time instants for side layout. (<b>a</b>) SedWaveFoam results; (<b>b</b>) FLOW-3D results.</p>
Full article ">Figure 11
<p>Bed elevation and suspended sediment concentration (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi>ϕ</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msup> <mo>)</mo> </mrow> </semantics></math> in the vicinity of the structure at various time instants for center layout. (<b>a</b>) SedWaveFoam results; (<b>b</b>) FLOW-3D results.</p>
Full article ">Figure 12
<p>Plan view of sandy bed at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s. Upper panels: side layout. Lower panels: center layout. (<b>a</b>,<b>e</b>) SedWaveFoam results; (<b>b</b>,<b>f</b>) FLOW-3D results; (<b>c</b>,<b>g</b>) difference between SedWaveFoam results and measurement; (<b>d</b>,<b>h</b>) difference between FLOW-3D results and measurement.</p>
Full article ">Figure 13
<p>Elevation of sandy bed along vortex trajectories at seaside and leeside of the structure at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s. Upper panels: side layout. Lower panels: center layout. Gray, red, and black lines represent measurement, SedWaveFoam, and FLOW-3D results, respectively.</p>
Full article ">Figure 14
<p>Temporal evolution of maximum scour depth at seaside and leeside of the structure. Upper panels: side layout. Lower panels: center layout. Red and black lines represent SedWaveFoam and FLOW-3D results, respectively. X represents the measured maximum scour depth.</p>
Full article ">Figure 15
<p>Cross-sectional view of sandy bed along line A–A and D–D at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> </mrow> </semantics></math> 15 s: (<b>a</b>) side layout; (<b>b</b>) center layout. Gray, red, and black solid lines represent measurement, SedWaveFoam, and FLOW-3D results, respectively. Gray dashed line represents the structure.</p>
Full article ">
17 pages, 3601 KiB  
Article
Simulation and Diagnosis of Physical Precipitation Process of Local Severe Convective Rainstorm in Ningbo
by Tingting Lu, Yeyi Ding, Zan Liu, Fan Wu, Guoqiang Xue, Chengming Zhang and Yuan Fu
Atmosphere 2024, 15(6), 658; https://doi.org/10.3390/atmos15060658 - 30 May 2024
Viewed by 608
Abstract
On 31 July 2021, Ningbo, an eastern coast city in China, experienced a severe convective rainstorm, characterized by intense short-duration precipitation extremes with a maximum rainfall rate of 130 mm h−1. In this research, we first analyzed this rainstorm using Doppler [...] Read more.
On 31 July 2021, Ningbo, an eastern coast city in China, experienced a severe convective rainstorm, characterized by intense short-duration precipitation extremes with a maximum rainfall rate of 130 mm h−1. In this research, we first analyzed this rainstorm using Doppler radar and precipitation observation and then conducted high-resolution simulation for it. A three-dimensional precipitation diagnostic equation is introduced to quantitatively analyze the microphysical processes during the rainstorm. It is shown that this rainstorm was triggered and developed locally in central Ningbo under favorable large-scale quasi-geostrophic conditions and local conditions. In the early stage, the precipitation increase is mainly driven by the strong convergence of water vapor, and a noticeable increase in both the intensity and spatial extent of uplift promotes the upward transportation of water vapor. As the water vapor flux and associated convergence weaken in the later stage, the precipitation reduces accordingly. Cloud microphysical processes are also important in the entire precipitation process. The early stage updraft supports the escalations in raindrops, with the notable fluctuations in raindrop concentrations directly linked to variations in ground precipitation intensity. The behavior of graupel particles is intricately connected to their melting as they fall below the zero-degree layer. Although cloud water and snow exhibit changes during this period, the magnitudes of these adjustments are considerably less pronounced than those in raindrops and graupels, highlighting the differentiated response of various condensates to the convective dynamics. These results can help deepen the understanding of local severe rainstorms and provide valuable scientific references for practical forecasting. Full article
(This article belongs to the Special Issue Characteristics of Extreme Climate Events over China)
Show Figures

Figure 1

Figure 1
<p>Model nested domains configuration.</p>
Full article ">Figure 2
<p>Distribution of the accumulated rainfall amount (shaded; unit: mm) during the rainstorm event for (<b>a</b>) observation from automatic weather stations and (<b>b</b>) simulated rainfall amount.</p>
Full article ">Figure 3
<p>The 500 hPa geopotential height (thick blue lines, unit: gpm) and 850 hPa wind fields (wind bar, unit: m s<sup>−1</sup>) at (<b>a</b>) 0800 BST 31 July, (<b>b</b>)1400 BST 31 July, (<b>c</b>) 2000 BST 31 July, and (<b>d</b>) 0200 BST 1 August 2021.</p>
Full article ">Figure 4
<p>The same as <a href="#atmosphere-15-00658-f003" class="html-fig">Figure 3</a>, but for the simulations (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 5
<p>The combined radar reflectivity observed in Ningbo (shaded, unit: dBZ) from 1500 BST to 2230 BST 31 July in 2021.</p>
Full article ">Figure 6
<p>Same as <a href="#atmosphere-15-00658-f005" class="html-fig">Figure 5</a>, but for the simulated radar reflectivity.</p>
Full article ">Figure 7
<p>Distribution of hourly rainfall in Ningbo (shaded, units: mm h<sup>−1</sup>) from 1500 BST 31 July to 1800 BST 31 July 2021. The left column shows the observations (<b>a1</b>–<b>d1</b>) and the right column shows the simulations (<b>a2</b>–<b>d2</b>).</p>
Full article ">Figure 8
<p>Temporal evolutions of the area-averaged (29.2°~29.5° N, 121.0°~121.4° E) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mi mathvariant="normal">S</mi> </mrow> </msub> </mrow> </semantics></math>, moisture-related processes (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi mathvariant="normal">W</mi> <mi mathvariant="normal">V</mi> </mrow> </msub> </mrow> </semantics></math>), change rates for hydrometeor-related processes (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">L</mi> </mrow> </msub> </mrow> </semantics></math>), and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi mathvariant="normal">C</mi> <mi mathvariant="normal">I</mi> </mrow> </msub> </mrow> </semantics></math> from 1500 BST to 1900 BST 31 July 2021.</p>
Full article ">Figure 9
<p>Area-averaged (29.2°~29.5° N, 121.0°~121.4° E) vertical profiles of hydrometeor mixing ratio (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math> for graupel, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> for snow, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math> for ice, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi>r</mi> </mrow> </msub> </mrow> </semantics></math> for raindrops, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Q</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> for cloud water, units: 10<sup>−3</sup> kg/kg; w for vertical speed, unit: m/s) from 1500 LST (notation in the sub-figures: 1500) to 1830 LST 31 July 2021.</p>
Full article ">
26 pages, 12365 KiB  
Article
Improving the Fuel Economy and Energy Efficiency of Train Cab Climate Systems, Considering Air Recirculation Modes
by Ivan Panfilov, Alexey N. Beskopylny and Besarion Meskhi
Energies 2024, 17(9), 2224; https://doi.org/10.3390/en17092224 - 5 May 2024
Cited by 1 | Viewed by 1392
Abstract
Current developments in vehicles have generated great interest in the research and optimization of heating, ventilation, and air conditioning (HVAC) systems as a factor to reduce fuel consumption. One of the key trends for finding solutions is the intensive development of electric transport [...] Read more.
Current developments in vehicles have generated great interest in the research and optimization of heating, ventilation, and air conditioning (HVAC) systems as a factor to reduce fuel consumption. One of the key trends for finding solutions is the intensive development of electric transport and, consequently, additional requirements for reducing energy consumption and modifying climate systems. Of particular interest is the optimal functioning of comfort and life support systems during air recirculation, i.e., when there is a complete or partial absence of outside air supply, in particular to reduce energy consumption or when the environment is polluted. This work examines numerical models of airfields (temperature, speed, and humidity) and also focuses on the concentration of carbon dioxide and oxygen in the cabin, which is a critical factor for ensuring the health of the driver and passengers. To build a mathematical model, the Navier–Stokes equations with energy, continuity, and diffusion equations are used to simulate the diffusion of gases and air humidity. In the Ansys Fluent finite volume analysis package, the model is solved numerically using averaged RANS equations and k-ω turbulence models. The cabin of a mainline locomotive with two drivers, taking into account their breathing, is used as a transport model. The problem was solved in a nonstationary formulation for the design scenario of summer and winter, the time of stabilization of the fields was found, and graphs were constructed for different points in time. A comparative analysis of the uniformity of fields along the height of the cabin was carried out with different locations of deflectors, and optimal configurations were found. Energy efficiency values of the climate system operation in recirculation operating modes were obtained. A qualitative assessment of the driver’s blowing directions under different circulation and recirculation modes is given from the point of view of the concentration of carbon dioxide in the breathing area. The proposed solution makes it possible to reduce electricity consumption from 3.1 kW to 0.6 kW and in winter mode from 11.6 kW to 3.9 kW and save up to 1.5 L/h of fuel. The conducted research can be used to develop modern energy-efficient and safe systems for providing comfortable climate conditions for drivers and passengers of various types of transport. Full article
(This article belongs to the Section B: Energy and Environment)
Show Figures

Figure 1

Figure 1
<p>2TE25KM—mainline freight two-section diesel locomotive with AC-DC electrical transmission: (<b>a</b>) general view; (<b>b</b>) control system.</p>
Full article ">Figure 2
<p>Typical scheme of cabin: (<b>a</b>) actual diagram; (<b>b</b>) 3D model.</p>
Full article ">Figure 3
<p>Geometry model for FEM analysis: (<b>a</b>) orange—inputs; and (<b>b</b>) blue—outputs.</p>
Full article ">Figure 4
<p>Mesh of finite volumes: (<b>a</b>) axonometric view; (<b>b</b>) side view.</p>
Full article ">Figure 5
<p>Temperature values versus time in summer mode.</p>
Full article ">Figure 6
<p>CO<sub>2</sub> concentration values over time in summer mode.</p>
Full article ">Figure 7
<p>CO<sub>2</sub> concentration values over time in summer mode on the indicated surfaces.</p>
Full article ">Figure 8
<p>CO<sub>2</sub> concentration values in the center over time in summer mode, as well as in recirculation mode.</p>
Full article ">Figure 9
<p>CO<sub>2</sub> concentration values over time in summer mode on the indicated surfaces, as well as for recirculation mode.</p>
Full article ">Figure 10
<p>Temperature values over time in winter mode.</p>
Full article ">Figure 11
<p>CO<sub>2</sub> concentration values over time in winter mode.</p>
Full article ">Figure 12
<p>CO<sub>2</sub> concentration values over time in winter mode on the indicated surfaces.</p>
Full article ">Figure 13
<p>Power of the climate system in circulation and recirculation mode.</p>
Full article ">Figure 14
<p>Temperature, °C: (<b>a</b>) summer mode; (<b>b</b>) winter mode.</p>
Full article ">Figure 15
<p>Temperature on the same numerical scale, °C: (<b>a</b>) summer mode; (<b>b</b>) winter mode.</p>
Full article ">Figure 16
<p>Velocity field, m/s: (<b>a</b>)—summer mode; (<b>b</b>)—winter mode.</p>
Full article ">Figure 17
<p>Relative humidity, %: (<b>a</b>)—summer mode; (<b>b</b>)—winter mode.</p>
Full article ">Figure 18
<p>Relative humidity on the same numerical scale, %: (<b>a</b>)—summer mode; (<b>b</b>)—winter mode.</p>
Full article ">Figure 19
<p>Gas concentration, volume fraction: (<b>a</b>)—summer mode; (<b>b</b>)—winter mode.</p>
Full article ">Figure 20
<p>Temperature measurement: (<b>a</b>)—Meteoskop-M [<a href="#B89-energies-17-02224" class="html-bibr">89</a>]; (<b>b</b>) blue and red—points for measuring temperature.</p>
Full article ">Figure 21
<p>Temperature measurement points: (<b>a</b>)—summer mode; (<b>b</b>)—winter mode.</p>
Full article ">Figure 22
<p>Temperatures on the horizontal axis OZ.</p>
Full article ">Figure 23
<p>Temperatures on the vertical axis OY.</p>
Full article ">Figure 24
<p>Velocities on the horizontal axis OZ.</p>
Full article ">Figure 25
<p>Velocities on the vertical axis OY.</p>
Full article ">Figure 26
<p>CO<sub>2</sub> streamlines for the summer mode for the case of the deflector located under the windshield: (<b>a</b>) frontal view; (<b>b</b>) side view.</p>
Full article ">Figure 27
<p>CO<sub>2</sub> streamlines for the summer mode for the case of deflectors located under the windshield and on the roof: (<b>a</b>) frontal view; (<b>b</b>) side view.</p>
Full article ">
22 pages, 34917 KiB  
Article
Unsteady Subsonic/Supersonic Flow Simulations in 3D Unstructured Grids over an Acoustic Cavity
by Guillermo Araya
Fluids 2024, 9(4), 92; https://doi.org/10.3390/fluids9040092 - 17 Apr 2024
Viewed by 1421
Abstract
In this study, the unsteady Reynolds-averaged Navier–Stokes (URANS) equations are employed in conjunction with the Menter Shear Stress Transport (SST)-Scale-Adaptive Simulation (SAS) turbulence model in compressible flow, with an unstructured mesh and complex geometry. While other scale-resolving approaches in space and time, such [...] Read more.
In this study, the unsteady Reynolds-averaged Navier–Stokes (URANS) equations are employed in conjunction with the Menter Shear Stress Transport (SST)-Scale-Adaptive Simulation (SAS) turbulence model in compressible flow, with an unstructured mesh and complex geometry. While other scale-resolving approaches in space and time, such as direct numerical simulation (DNS) and large-eddy simulation (LES), supply more comprehensive information about the turbulent energy spectrum of the fluctuating component of the flow, they imply computationally intensive situations, usually performed over structured meshes and relatively simple geometries. In contrast, the SAS approach is designed according to “physically” prescribed length scales of the flow. More precisely, it operates by locally comparing the length scale of the modeled turbulence to the von Karman length scale (which depends on the local first- and second fluid velocity derivatives). This length-scale ratio allows the flow to dynamically adjust the local eddy viscosity in order to better capture the large-scale motions (LSMs) in unsteady regions of URANS simulations. While SAS may be constrained to model only low flow frequencies or wavenumbers (i.e., LSM), its versatility and low computational cost make it attractive for obtaining a quick first insight of the flow physics, particularly in those situations dominated by strong flow unsteadiness. The selected numerical application is the well-known M219 three-dimensional rectangular acoustic cavity from the literature at two different free-stream Mach numbers, M (0.85 and 1.35) and a length-to-depth ratio of 5:1. Thus, we consider the “deep configuration” in experiments by Henshaw. The SST-SAS model demonstrates a satisfactory compromise between simplicity, accuracy, and flow physics description. Full article
Show Figures

Figure 1

Figure 1
<p>Experiment schematic and dimensions adapted from [<a href="#B15-fluids-09-00092" class="html-bibr">15</a>]: (<b>a</b>) right-side view, (<b>b</b>) top view, and (<b>c</b>) front view (flow from left to right in (<b>a</b>,<b>b</b>)).</p>
Full article ">Figure 2
<p>RMS of pressure fluctuations at the deep cavity ceiling and rig centerline (<span class="html-italic">y</span> = 0); from [<a href="#B15-fluids-09-00092" class="html-bibr">15</a>].</p>
Full article ">Figure 3
<p>Several schematics of the computational box and hybrid mesh: (<b>a</b>) lateral view with dimensions and boundary conditions (flow from left to right), (<b>b</b>) isometric view and half-plane of the cavity (flow from left to right), and (<b>c</b>) interior view (from upstream of the cavity).</p>
Full article ">Figure 3 Cont.
<p>Several schematics of the computational box and hybrid mesh: (<b>a</b>) lateral view with dimensions and boundary conditions (flow from left to right), (<b>b</b>) isometric view and half-plane of the cavity (flow from left to right), and (<b>c</b>) interior view (from upstream of the cavity).</p>
Full article ">Figure 4
<p>Time variation of the total drag in the acoustic cavity at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 5
<p>Root mean square of pressure fluctuations in the acoustic cavity ceiling at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85 and 1.35: coarse and fine mesh results.</p>
Full article ">Figure 6
<p>Root mean square of pressure fluctuations in the acoustic cavity ceiling: comparison with experimental values at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85 and 1.35.</p>
Full article ">Figure 7
<p>Root mean square of pressure fluctuations in the acoustic cavity at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 8
<p>Root mean square of pressure fluctuations in the acoustic cavity at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 1.35.</p>
Full article ">Figure 9
<p>Contours of the turbulence length scale over and inside the cavity: (<b>a</b>) longitudinal plane and (<b>b</b>) cross-sectional planes at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 9 Cont.
<p>Contours of the turbulence length scale over and inside the cavity: (<b>a</b>) longitudinal plane and (<b>b</b>) cross-sectional planes at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 10
<p>Iso-contours of instantaneous spanwise vorticity in 1/s (flow from left to right) at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> </mrow> </semantics></math> 0 and <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 11
<p>Iso-contours of instantaneous streamwise velocity normalized by the free-stream velocity (flow from left to right) at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> </mrow> </semantics></math> 0 and <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 12
<p>Iso-surfaces of <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>−</mo> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> <mi>e</mi> <mi>r</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </semantics></math>, positive values in red (vortex cores), negative values in blue (highly deformed flow regions), at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 0.85.</p>
Full article ">Figure 13
<p>Types of supersonic cavity flows.</p>
Full article ">Figure 13 Cont.
<p>Types of supersonic cavity flows.</p>
Full article ">Figure 14
<p>Iso-surfaces of <math display="inline"><semantics> <mrow> <mi>Q</mi> <mo>−</mo> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> <mi>e</mi> <mi>r</mi> <mi>i</mi> <mi>o</mi> <mi>n</mi> </mrow> </semantics></math>, positive values in red (vortex cores), negative values in blue (highly deformed flow regions), at <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 1.35.</p>
Full article ">Figure 15
<p>Iso-contours of instantaneous spanwise vorticity in 1/s (flow from left to right) at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> </mrow> </semantics></math> 0 and <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 1.35.</p>
Full article ">Figure 16
<p>Iso-contours of instantaneous spanwise vorticity in 1/s (flow from left to right) at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> </mrow> </semantics></math> 0.25 and <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 1.35.</p>
Full article ">Figure 17
<p>Iso-contours of instantaneous streamwise velocity normalized by the free-stream velocity (flow from left to right) at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> </mrow> </semantics></math> 0 and <math display="inline"><semantics> <msub> <mi>M</mi> <mo>∞</mo> </msub> </semantics></math> = 1.35.</p>
Full article ">
17 pages, 6648 KiB  
Article
Responses of Soil Phosphorus Cycling-Related Microbial Genes to Thinning Intensity in Cunninghamia lanceolata Plantations
by Dongxu Ma, Jiaqi Wang, Kuaiming Chen, Weili Lan, Yiquan Ye, Xiangqing Ma and Kaimin Lin
Forests 2024, 15(3), 440; https://doi.org/10.3390/f15030440 - 26 Feb 2024
Cited by 1 | Viewed by 1592
Abstract
Background: Microorganisms are important regulators of soil phosphorus cycling and phosphorus availability in Chinese fir (Cunninghamia lanceolata (Lamb.) Hook) plantations. However, the effects of thinning on soil phosphorus cycling by microbes in C. lanceolata plantations remain unclear. Methods: We performed a metagenomic [...] Read more.
Background: Microorganisms are important regulators of soil phosphorus cycling and phosphorus availability in Chinese fir (Cunninghamia lanceolata (Lamb.) Hook) plantations. However, the effects of thinning on soil phosphorus cycling by microbes in C. lanceolata plantations remain unclear. Methods: We performed a metagenomic sequencing analysis to investigate how thinning intensities (weak, moderate, and heavy) alter phosphorus cycling related microbial genes and their regulatory effects on soil phosphorus availability in C. lanceolata plantations. Results: Following heavy thinning, the contents of available and labile phosphorus increased by 13.8% and 36.9%, respectively, compared to moderate and weak thinning. Moreover, the relative abundance of genes associated with inorganic phosphorus solubilization increased significantly with the increase in thinning intensity, whereas genes associated with phosphorus uptake and transport significantly decreased. The metagenomic analysis results indicate that Acidobacteria (47.6%–53.5%), Proteobacteria (17.9%–19.1%), and Actinobacteria (11.7%–12.8%) are the major contributors to the functional phosphorus cycling genes in the soil. The random forest analysis results suggested that gcd, plc, phoN, ugpA, and phoR were the critical genes involved in the transformation and use of phosphorus, which in turn increased soil phosphorus availability. Structural equation modeling revealed that soil pH was the primary factor influencing changes in functional genes associated with phosphorus cycling in C. lanceolata plantations. Specifically, soil pH (ranging from 4.3 to 4.9) were positively correlated with genes involved in inorganic phosphate solubilization and organic phosphate mineralization, while negatively correlated with genes related to phosphorus uptake and transport. Conclusions: Taken together, our results demonstrate that the enhanced microbe-mediated mineralization of organic phosphorus and solubilization of inorganic phosphorus are suppressed when uptake and transportation are the mechanisms responsible for the increased soil phosphorus availability under appropriate thinning intensities. Changes in the soil microbial community and phosphorus cycling genes in response to different thinning intensities may maintain soil functionality and nutrient balance in C. lanceolata plantations. These findings contribute to a better understanding of the mechanisms underlying the microbial mediation of phosphorus cycling in the soil of C. lanceolata plantations. Full article
(This article belongs to the Special Issue Adaptive Mechanisms of Tree Seedlings to Adapt to Stress)
Show Figures

Figure 1

Figure 1
<p>Non-metric multidimensional scaling (based on the Bray–Curtis distance) of phosphorus cycling functional genes in response to different thinning intensities. The solid line represents the 95% confidence interval.</p>
Full article ">Figure 2
<p>Differences in the relative abundances of phosphorus cycling-related functional genes in response to different thinning intensities (mean ± standard deviation; n = 3). The genes were classified according to their functions. (<b>a</b>) Four phosphorus cycling categories. (<b>b</b>) Organic acid formation. (<b>c</b>) Fifteen phosphorus cycling functional groups. WT: weak thinning; MT: moderate thinning; HT: moderate thinning. * and ** represent significant correlations at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">Figure 3
<p>Contributions of the soil microorganisms involved in phosphorus cycling to the relative abundances of 23 key phosphorus cycling genes related to (<b>a</b>) inorganic phosphorus solubilization, (<b>b</b>) phosphorus transport, (<b>c</b>) phosphorus starvation response regulation, and (<b>d</b>) organic phosphorus mineralization. WT: weak thinning; MT: moderate thinning; HT: moderate thinning.</p>
Full article ">Figure 4
<p>Heatmap of the correlations between soil properties, phosphorus status, phosphorus fractions, and the genes involved in soil phosphorus cycling. * and ** represent significant correlations at <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">Figure 5
<p>Structural equation model presenting the direct and indirect effects of thinning on soil phosphorus availability. The width of the solid lines is directly proportional to the significance of the relationship between variables, with solid and dashed lines representing positive and negative correlations, respectively. CMIN/DF, chi-square minimum/degrees of freedom; TLI, Tucker–Lewis index; CFI, comparative fit index; RMSEA, root mean square error of approximation. *, **, and *** represent significant correlations at <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001, respectively.</p>
Full article ">
Back to TopTop