[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (207)

Search Parameters:
Keywords = cooling air velocity

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
28 pages, 2842 KiB  
Review
Heat Transfer Performance Factors in a Vertical Ground Heat Exchanger for a Geothermal Heat Pump System
by Khaled Salhein, C. J. Kobus, Mohamed Zohdy, Ahmed M. Annekaa, Edrees Yahya Alhawsawi and Sabriya Alghennai Salheen
Energies 2024, 17(19), 5003; https://doi.org/10.3390/en17195003 - 8 Oct 2024
Viewed by 712
Abstract
Ground heat pump systems (GHPSs) are esteemed for their high efficiency within renewable energy technologies, providing effective solutions for heating and cooling requirements. These GHPSs operate by utilizing the relatively constant temperature of the Earth’s subsurface as a thermal source or sink. This [...] Read more.
Ground heat pump systems (GHPSs) are esteemed for their high efficiency within renewable energy technologies, providing effective solutions for heating and cooling requirements. These GHPSs operate by utilizing the relatively constant temperature of the Earth’s subsurface as a thermal source or sink. This feature allows them to perform greater energy transfer than traditional heating and cooling systems (i.e., heating, ventilation, and air conditioning (HVAC)). The GHPSs represent a sustainable and cost-effective temperature-regulating solution in diverse applications. The ground heat exchanger (GHE) technology is well known, with extensive research and development conducted in recent decades significantly advancing its applications. Improving GHE performance factors is vital for enhancing heat transfer efficiency and overall GHPS performance. Therefore, this paper provides a comprehensive review of research on various factors affecting GHE performance, such as soil thermal properties, backfill material properties, borehole depth, spacing, U-tube pipe properties, and heat carrier fluid type and velocity. It also discusses their impact on heat transfer efficiency and proposes optimal solutions for improving GHE performance. Full article
(This article belongs to the Special Issue Advances in Refrigeration and Heat Pump Technologies)
Show Figures

Figure 1

Figure 1
<p>Installed capacity (MWt) of geothermal heat pump systems worldwide from 1995 to 2020 [<a href="#B18-energies-17-05003" class="html-bibr">18</a>].</p>
Full article ">Figure 2
<p>Schematic diagram of geothermal heat pump system [<a href="#B7-energies-17-05003" class="html-bibr">7</a>].</p>
Full article ">Figure 3
<p>Schematic diagram of single U-tube vertical ground heat exchanger.</p>
Full article ">Figure 4
<p>Thermal performance factors of the vertical ground heat exchanger [<a href="#B7-energies-17-05003" class="html-bibr">7</a>].</p>
Full article ">Figure 5
<p>Relationship between the geothermal heat transfer rate and thermal resistance [<a href="#B5-energies-17-05003" class="html-bibr">5</a>].</p>
Full article ">Figure 6
<p>Relationship between the geothermal heat transfer rate and coefficient of performance [<a href="#B5-energies-17-05003" class="html-bibr">5</a>].</p>
Full article ">Figure 7
<p>Shank space in a horizontal cross-section of a U-tube pipe in a vertical ground heat exchanger (GHE) [<a href="#B5-energies-17-05003" class="html-bibr">5</a>].</p>
Full article ">Figure 8
<p>Relationship between the water velocity and geothermal pipe length [<a href="#B5-energies-17-05003" class="html-bibr">5</a>].</p>
Full article ">Figure 9
<p>Illustration of the behavior of water temperature inside a vertical single U-tube pipe at various velocities (0.35, 0.45, 0.9, and 1.2 m/s) during heating mode (i.e., winter operation) [<a href="#B7-energies-17-05003" class="html-bibr">7</a>].</p>
Full article ">Figure 10
<p>Illustration of the behavior of water temperature inside a vertical single U-tube pipe at various velocities (0.35, 0.45, 0.9, and 1.2 m/s) during cooling mode (i.e., summer operation) [<a href="#B7-energies-17-05003" class="html-bibr">7</a>].</p>
Full article ">
17 pages, 6892 KiB  
Article
Effect of Spray Characteristic Parameters on Friction Coefficient of Ultra-High-Strength Steel against Cemented Carbide
by Bangfu Wu, Minxiu Zhang, Biao Zhao, Benkai Li and Wenfeng Ding
Materials 2024, 17(19), 4867; https://doi.org/10.3390/ma17194867 - 3 Oct 2024
Viewed by 520
Abstract
Ultra-high-strength steels have been considered an essential material for aviation components owing to their excellent mechanical properties and superior fatigue resistance. When machining these steels, severe tool wear frequently results in poor surface quality and low machining efficiency, which is intimately linked to [...] Read more.
Ultra-high-strength steels have been considered an essential material for aviation components owing to their excellent mechanical properties and superior fatigue resistance. When machining these steels, severe tool wear frequently results in poor surface quality and low machining efficiency, which is intimately linked to the friction behavior at the tool–workpiece interface. To enhance the service life of tools, the adoption of efficient cooling methods is paramount. However, the understanding of friction behavior at the tool–workpiece interface under varying cooling conditions remains limited. In this work, both air atomization of cutting fluid (AACF) and ultrasonic atomization of cutting fluid (UACF) were employed, and their spray characteristic parameters, including droplet size distribution, droplet number density, and droplet velocity, were evaluated under different air pressures. Discontinuous sliding tests were conducted on the ultra-high-strength steel against cemented carbide and the effect of spray characteristic parameters on the adhesion friction coefficient was studied. The results reveal that ultrasonic atomization significantly improved the uniformity of droplet size distribution. An increase in air pressure resulted in an increase in both droplet number density and droplet velocity under both AACF and UACF conditions. Furthermore, the thickness of the liquid film was strongly dependent on the spray characteristic parameters. Additionally, UACF exhibited a reduction of 4.7% to 9.8% in adhesion friction coefficient compared to AACF. UACF provided the appropriate combination of spray characteristic parameters, causing an increased thickness of the liquid film, which subsequently exerted a positive impact on reducing the adhesion friction coefficient. Full article
(This article belongs to the Special Issue Cutting Processes for Materials in Manufacturing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of the cooling system for AACF and UACF conditions.</p>
Full article ">Figure 2
<p>Experimental setup for droplet measurement: (<b>a</b>) droplet deposition device and (<b>b</b>) optical microscope.</p>
Full article ">Figure 3
<p>The recognition process of droplet image: (<b>a</b>) original image, (<b>b</b>) binary image, and (<b>c</b>) statistics results of droplet size distribution.</p>
Full article ">Figure 4
<p>Experimental setup for the discontinuous sliding test under different cooling conditions: (<b>a</b>) schematic diagram of the discontinuous sliding process, (<b>b</b>) open ball–disc tribometer system, (<b>c</b>) holder, and (<b>d</b>) force measurement system and cooling system.</p>
Full article ">Figure 5
<p>WC cemented carbide ball: (<b>a</b>) SEM image of microstructure and (<b>b</b>) EDS mapping.</p>
Full article ">Figure 6
<p>Ultra-high-strength steel: (<b>a</b>) SEM image of microstructure and (<b>b</b>) EDS mapping.</p>
Full article ">Figure 7
<p>Discontinuous sliding process: (<b>a</b>) force transformation relationship under maximum sliding depth and (<b>b</b>) projection area of the contact surface in a local coordinate system.</p>
Full article ">Figure 8
<p>Calculation process of adhesion friction coefficient.</p>
Full article ">Figure 9
<p>Histogram of droplet size distribution under different air pressures: (<b>a</b>) <span class="html-italic">P</span> = 140 kPa, (<b>b</b>) <span class="html-italic">P</span> = 180 kPa, (<b>c</b>) <span class="html-italic">P</span> = 220 kPa, (<b>d</b>) <span class="html-italic">P</span> = 260 kPa, and (<b>e</b>) <span class="html-italic">P</span> = 300 kPa.</p>
Full article ">Figure 10
<p>Statistical results of droplet size distribution and number density under different air pressures: (<b>a</b>) average droplet diameter <span class="html-italic">d</span><sub>a</sub>, (<b>b</b>) standard deviation of droplet diameter <span class="html-italic">σ</span><sub>d</sub>, and (<b>c</b>) droplet number density <span class="html-italic">N</span><sub>d</sub>.</p>
Full article ">Figure 11
<p>Droplet velocity under different air pressures: (<b>a</b>) variation of droplet velocity with spray distance under AACF, (<b>b</b>) variation of droplet velocity with spray distance under UACF, and (<b>c</b>) droplet velocity at a spray distance of 55 mm.</p>
Full article ">Figure 12
<p>Effect of air pressure on the dimensionless number under AACF and UACF.</p>
Full article ">Figure 13
<p>Effect of air pressure on the adhesion friction coefficient under different cooling conditions: (<b>a</b>) adhesion friction coefficient curve under AACF, (<b>b</b>) adhesion friction coefficient curve under UACF, and (<b>c</b>) variation of average adhesion friction coefficient with air pressure in the stable stage.</p>
Full article ">Figure 14
<p>Effect of air pressure on the worn surface of the WC ball under (<b>a</b>–<b>e</b>) 140 kPa, 180 kPa, 220 kPa, 260 kPa, and 300 kPa for AACF, (<b>f</b>–<b>j</b>) 140 kPa, 180 kPa, 220 kPa, 260 kPa, and 300 kPa for UACF.</p>
Full article ">
23 pages, 5482 KiB  
Article
Developing a Chained Simulation Method for Quantifying Cooling Energy in Buildings Affected by the Microclimate of Avenue Trees
by Bryon Flowers and Kuo-Tsang Huang
Atmosphere 2024, 15(10), 1150; https://doi.org/10.3390/atmos15101150 - 25 Sep 2024
Viewed by 401
Abstract
This paper introduces a methodology aimed at bridging the gap between building energy simulation and urban climate modeling. A coupling method was developed through the Building Control Virtual Test Bed (BCVTB) and applied to a case study in Taipei City, Taiwan, to address [...] Read more.
This paper introduces a methodology aimed at bridging the gap between building energy simulation and urban climate modeling. A coupling method was developed through the Building Control Virtual Test Bed (BCVTB) and applied to a case study in Taipei City, Taiwan, to address the microclimate factors of street trees crucial to cooling energy consumption. The use of the Urban Weather Generator for weather file modification revealed a 0.63 °C average air temperature disparity. The coupling method emphasized the importance of accurate wind speed and convective heat transfer coefficients (CHTCs) on building surfaces in determining cooling energy. The results indicated that elevated CHTC values amplify heat exchange, with higher wind velocities playing a crucial role in heat dissipation. The presence of street trees was found to significantly reduce heat flux penetration, leading to a reduction in building surface temperatures by as much as 9.5% during hot months. The cooling energy was lowered by 16.7% in the BCVTB simulations that included trees compared to those without trees. The EnergyPlus-only simulations underestimated the cooling energy needs by approximately 9.3% during summer months. This research offers valuable insights into the complex interactions between buildings and their environments. The results highlight the importance of trees and shading in mitigating the heat island effect and improving energy-efficient urban planning. Full article
(This article belongs to the Section Biometeorology and Bioclimatology)
Show Figures

Figure 1

Figure 1
<p>Overview of research methodology.</p>
Full article ">Figure 2
<p>Case study building model: (<b>a</b>) the 3D building is enclosed by structures on all sides that represent other buildings or surfaces in EnergyPlus; (<b>b</b>) the building model in ENVI-met: the location of the target building that is surrounded by adjacent buildings of a similar structure and trees that are equidistant.</p>
Full article ">Figure 3
<p>The difference between the dry-bulb temperature profile of the TMY3 (avg. T of 23.33 °C) and the dry-bulb temperature profile of the UWG-produced weather file (avg. T of 23.96 °C).</p>
Full article ">Figure 4
<p>ENVI-met average wind speed results for north, south, east, and west façades of building model and the wind speed results of the standard TMY3 for Taipei.</p>
Full article ">Figure 5
<p>Summary of ENVI-met ambient air temperature results: the line in the plot is the average temperature that was calculated by averaging the hottest and coldest day values; the hottest and coldest days are also shown as the upper and lower lines, respectively.</p>
Full article ">Figure 6
<p>BCVTB surface temperature results for spring (<b>top left</b>), summer (<b>top right</b>), autumn (<b>bottom left</b>), and winter (<b>bottom right</b>) months. The ribbon plots show the upper and lower bounds of surface temperature results, which were carried out for the hottest and coldest days of each month.</p>
Full article ">Figure 7
<p>Average CHTC results from coldest and hottest days of April to September months from EnergyPlus simulations and BCVTB-coupled simulations.</p>
Full article ">Figure 8
<p>Violin plots of days that contained the hottest surface temperatures of the year. The violin plots are also density plots that represents the frequency of surface temperature values on the y-axis. The red dots in the boxplot are the median surface temperature values. Colored dots represent the CHTC values and the corresponding surface temperature for that timestep.</p>
Full article ">Figure 9
<p>BCVTB surface temperature comparison between surface temperatures from simulations with trees and without trees for May: (<b>a</b>) daytime surface temperature variation for May; (<b>b</b>) nighttime surface temperature variation.</p>
Full article ">Figure 10
<p>Surface temperature difference (ΔT) between BCVTB simulations with trees and without trees for months that require cooling energy.</p>
Full article ">Figure 11
<p>Total cooling energy consumption comparison between BCVTB-coupled simulation with trees, BCVTB-coupled simulation without trees, and EnergyPlus-only simulation cooling energy.</p>
Full article ">
17 pages, 5959 KiB  
Article
Effects of Different Cooling Treatments on Heated Granite: Insights from the Physical and Mechanical Characteristics
by Qinming Liang, Gun Huang, Jinyong Huang, Jie Zheng, Yueshun Wang and Qiang Cheng
Materials 2024, 17(18), 4539; https://doi.org/10.3390/ma17184539 - 15 Sep 2024
Viewed by 521
Abstract
The exploration of Hot Dry Rock (HDR) geothermal energy is essential to fulfill the energy demands of the increasing population. Investigating the physical and mechanical properties of heated rock under different cooling methods has significant implications for the exploitation of HDR. In this [...] Read more.
The exploration of Hot Dry Rock (HDR) geothermal energy is essential to fulfill the energy demands of the increasing population. Investigating the physical and mechanical properties of heated rock under different cooling methods has significant implications for the exploitation of HDR. In this study, ultrasonic testing, uniaxial strength compression experiments, Brazilian splitting tests, nuclear magnetic resonance (NMR), and scanning electron microscope (SEM) were conducted on heated granite after different cooling methods, including cooling in air, cooling in water, cooling in liquid nitrogen, and cycle cooling in liquid nitrogen. The results demonstrated that the density, P-wave velocity (Vp), uniaxial compressive strength (UCS), tensile strength (σt), and elastic modulus (E) of heated granite tend to decrease as the cooling rate increases. Notably, heated granite subjected to cyclic liquid nitrogen cooling exhibits a more pronounced decline in physical and mechanical properties and a higher degree of damage. Furthermore, the cooling treatments also lead to an increase in rock pore size and porosity. At a faster cooling rate, the fracture surfaces of the granite transition from smooth to rough, suggesting enhanced fracture propagation and complexity. These findings provide critical theoretical insights into optimizing stimulation performance strategies for HDR exploitation. Full article
(This article belongs to the Special Issue Manufacturing, Characterization and Modeling of Advanced Materials)
Show Figures

Figure 1

Figure 1
<p>Samples and equipment. (<b>a</b>) Specimens for uniaxial compression strength experiment and Brazilian split tests. (<b>b</b>) MTS815 experimental apparatus. (<b>c</b>) AG-250kN IS Electronic Precision Material Testing Machine.</p>
Full article ">Figure 2
<p>Changes in mass loss rate and volume expansion rate of granite under different heating temperatures and cooling treatments. (<b>a</b>) Results of the temperature at 200 °C. (<b>b</b>) Results of the temperature at 300 °C.</p>
Full article ">Figure 3
<p>Density changes of granite after different cooling methods. (<b>a</b>) Results of the temperature at 200 °C. (<b>b</b>) Results of the temperature at 300 °C.</p>
Full article ">Figure 4
<p>Changes of <span class="html-italic">V</span><sub>p</sub> of granite treated with different cooling methods. (<b>a</b>) Results of the temperature at 200 °C. (<b>b</b>) Results of the temperature at 300 °C.</p>
Full article ">Figure 5
<p>Uniaxial compressive strength of granite after different cooling treatments.</p>
Full article ">Figure 6
<p>Changes in the tensile strength of granite under different cooling methods.</p>
Full article ">Figure 7
<p>Changes of elastic modulus of granite under different cooling methods.</p>
Full article ">Figure 8
<p>Pore size distribution of heated granite samples under different cooling methods. (<b>a</b>) Results of the temperature at 200 °C. (<b>b</b>) Results of the temperature at 300 °C.</p>
Full article ">Figure 9
<p>Changes in granite porosity under different cooling methods.</p>
Full article ">Figure 10
<p>Morphology of fracture surfaces of heated granite under different cooling treatments at 200 °C.</p>
Full article ">Figure 11
<p>Morphology of fracture surfaces of heated granite under different cooling treatments at 300 °C.</p>
Full article ">Figure 12
<p>Changes in <span class="html-italic">D</span> of heated granite under different cooling methods.</p>
Full article ">
24 pages, 5293 KiB  
Article
Computational Fluid Dynamics Study on Bottom-Hole Multiphase Flow Fields Formed by Polycrystalline Diamond Compact Drill Bits in Foam Drilling
by Lihong Wei and Jaime Honra
Fluids 2024, 9(9), 211; https://doi.org/10.3390/fluids9090211 - 10 Sep 2024
Viewed by 471
Abstract
High-temperature geothermal wells frequently employ foam drilling fluids and Polycrystalline Diamond Compact (PDC) drill bits. Understanding the bottom-hole flow field of PDC drill bits in foam drilling is essential for accurately analyzing their hydraulic structure design. Based on computational fluid dynamics (CFD) and [...] Read more.
High-temperature geothermal wells frequently employ foam drilling fluids and Polycrystalline Diamond Compact (PDC) drill bits. Understanding the bottom-hole flow field of PDC drill bits in foam drilling is essential for accurately analyzing their hydraulic structure design. Based on computational fluid dynamics (CFD) and multiphase flow theory, this paper establishes a numerical simulation technique for gas-liquid-solid multiphase flow in foam drilling with PDC drill bits, combined with a qualitative and quantitative hydraulic structure evaluation method. This method is applied to simulate the bottom-hole flow field of a six-blade PDC drill bit. The results show that the flow velocity of the air phase in foam drilling fluid is generally higher than that of the water phase. Some blades’ cutting teeth exhibit poor cleaning and cooling effects, with individual cutting teeth showing signs of erosion damage and cuttings cross-flow between channels. To address these issues, optimizing the nozzle spray angle and channel design is necessary to improve hydraulic energy distribution, enhance drilling efficiency, and extend drill bit life. This study provides new ideas and methods for developing geothermal drilling technology in the numerical simulation of a gas-liquid-solid three-phase flow field. Additionally, the combined qualitative and quantitative evaluation method offers new insights and approaches for research and practice in drilling engineering. Full article
(This article belongs to the Special Issue Multiphase Flow and Granular Mechanics)
Show Figures

Figure 1

Figure 1
<p>3D model of 6-blade PDC drill bit.</p>
Full article ">Figure 2
<p>Schematic diagram of the model of the bottom hole flow field calculation domain: (<b>a</b>) is the plan view; (<b>b</b>) is the bottom view.</p>
Full article ">Figure 3
<p>Bottom-hole flow field grid.</p>
Full article ">Figure 4
<p>Bottom-hole flow velocity contour: (<b>a</b>) is the velocity of air; (<b>b</b>) is the velocity of water.</p>
Full article ">Figure 5
<p>Surface velocity distribution of blade No. 1.</p>
Full article ">Figure 6
<p>Surface velocity distribution of blade No. 2.</p>
Full article ">Figure 7
<p>Surface velocity distribution of blade No. 3.</p>
Full article ">Figure 8
<p>Surface velocity distribution of blade No. 4.</p>
Full article ">Figure 9
<p>Surface velocity distribution of blade No.5.</p>
Full article ">Figure 10
<p>Surface velocity distribution of blade No. 6.</p>
Full article ">Figure 11
<p>Matching Diagram of Flow Channel Discharge Cuttings.</p>
Full article ">Figure 12
<p>Total Erosion Rate of Blade No. 1.</p>
Full article ">Figure 13
<p>Total Erosion Rate of Blade No. 2.</p>
Full article ">Figure 14
<p>Total Erosion Rate of Blade No. 3.</p>
Full article ">Figure 15
<p>Total Erosion Rate of Blade No. 4.</p>
Full article ">Figure 16
<p>Total Erosion Rate of Blade No.5.</p>
Full article ">Figure 17
<p>Total Erosion Rate of Blade No.6.</p>
Full article ">
13 pages, 6251 KiB  
Article
Using the Groundwater Cooling System and Phenolic Aldehyde Isolation Layer on Building Walls to Evaluation of Heat Effect
by Ting-Yu Chen and Wen-Pei Sung
Buildings 2024, 14(9), 2848; https://doi.org/10.3390/buildings14092848 - 10 Sep 2024
Viewed by 362
Abstract
This study examines the thermal performance of building walls under full sunlight conditions using various insulation strategies. Specifically, it evaluates: (1) the effects of heat on building walls and indoor spaces; (2) the impact of groundwater cooling systems on thermal environments; (3) the [...] Read more.
This study examines the thermal performance of building walls under full sunlight conditions using various insulation strategies. Specifically, it evaluates: (1) the effects of heat on building walls and indoor spaces; (2) the impact of groundwater cooling systems on thermal environments; (3) the influence of phenolic aldehyde insulation layers on heat transfer; and (4) the combined effects of groundwater cooling and phenolic aldehyde thermal insulation. Fluent–CFD (Computational Fluid Dynamics) was used in the study to simulate temperature transmission between the sun, the groundwater cooling system, and both indoor and outdoor spaces. Experimental analysis and simulations reveal that both the phenolic aldehyde insulation layer and the groundwater cooling system effectively reduce heat transfer, with the groundwater cooling system demonstrating the most significant impact. The phenolic aldehyde layer decreases the temperature difference between inner and outer walls by approximately 8 °C. The groundwater cooling system further reduces both inner and outer wall temperatures, helping to maintain cooler indoor environments. Simulation results indicate that, while the phenolic aldehyde layer effectively prevents external heat from penetrating into the room, it does not eliminate heat accumulation. In contrast, the groundwater cooling system efficiently dissipates heat, mitigating this issue. Groundwater analysis shows that maximum temperature differences occur at specific times of the day, with water flow effectively cooling the space. The combined use of the phenolic aldehyde insulation layer and the groundwater cooling system offers superior thermal performance. The phenolic layer provides effective heat blocking, while the groundwater system facilitates heat dissipation, optimizing indoor temperature and reducing air conditioning loads. This combination enhances overall comfort and energy efficiency, with the groundwater cooling system benefiting from reduced flow velocity and lower energy consumption. Full article
(This article belongs to the Collection Sustainable Buildings in the Built Environment)
Show Figures

Figure 1

Figure 1
<p>The experimental houses and reference point.</p>
Full article ">Figure 2
<p>The grid test uses different cell numbers.</p>
Full article ">Figure 3
<p>The building grid construction.</p>
Full article ">Figure 4
<p>The building grid construction. The cross-section temperature of the buildings in the original conditions and the temperature of the interior and exterior walls.</p>
Full article ">Figure 5
<p>The curve of temperature on the inner wall of east and west walls.</p>
Full article ">Figure 6
<p>The temperature of the cross-section of the cooling system and the temperature of the Interior and exterior walls on the east side.</p>
Full article ">Figure 7
<p>Temperature curves of the inner walls during the cooling pipeline test.</p>
Full article ">Figure 8
<p>Temperature curves of the outside walls during the cooling pipeline test.</p>
Full article ">Figure 9
<p>Temperature variations of the cooling pipeline.</p>
Full article ">Figure 10
<p>The temperature variation of the groundwater cooling test between the inflow and outflow.</p>
Full article ">Figure 11
<p>The temperature of the cross-section of the wall with phenolic layer and the temperature of the Interior and exterior walls on the east side.</p>
Full article ">Figure 12
<p>Temperature curves of the inner walls during the test with insulation materials.</p>
Full article ">Figure 13
<p>Temperature curves of the outside walls during the test with insulation materials.</p>
Full article ">Figure 14
<p>The temperature of the cross-section of the wall with two systems.</p>
Full article ">Figure 15
<p>Temperature variations of the east and west inner walls in the integration test.</p>
Full article ">Figure 16
<p>The curve of temperature on the outside wall of east and west.</p>
Full article ">Figure 17
<p>Temperature variations of the cooling pipeline in the integration test.</p>
Full article ">Figure 18
<p>The temperature variation of the integration test between the inflow and outflow.</p>
Full article ">
21 pages, 8901 KiB  
Article
CFD Investigation on Combined Ventilation System for Multilayer-Caged-Laying Hen Houses
by Changzeng Hu, Lihua Li, Yuchen Jia, Zongkui Xie, Yao Yu and Limin Huo
Animals 2024, 14(17), 2623; https://doi.org/10.3390/ani14172623 - 9 Sep 2024
Viewed by 425
Abstract
Mechanical ventilation is an important means of environmental control in multitier laying hen cages. The mainstream ventilation mode currently in use, negative-pressure ventilation (NPV), has the drawbacks of a large temperature difference before and after adjustment and uneven air velocity distribution. To solve [...] Read more.
Mechanical ventilation is an important means of environmental control in multitier laying hen cages. The mainstream ventilation mode currently in use, negative-pressure ventilation (NPV), has the drawbacks of a large temperature difference before and after adjustment and uneven air velocity distribution. To solve these problems, this study designed and analyzed a combined positive and negative-pressure ventilation system for laying hen cages. According to the principle of the conservation of mass to increase the inlet flow in the negative-pressure ventilation system on the basis of the addition of the pressure-wind body-built positive-and-negative-pressure-combined ventilation (PNCV) system, further, computational fluid dynamics (CFD) simulation was performed to analyze the distribution of environmental parameters in the chicken cage zone (CZ) with inlet angles of positive-pressure fans set at 45°, 90°, and 30°. Simulation results showed that the PNCV system increased the average air velocity in the CZ from 0.94 m/s to 1.04 m/s, 1.28 m/s, and 0.99 m/s by actively blowing air into the cage. The maximum temperature difference in the CZ with the PNCV system was 2.91 °C, 1.80 °C, and 3.78 °C, which were all lower than 4.46 °C, the maximum temperature difference in the CZ with the NPV system. Moreover, the relative humidity remained below 80% for the PNCV system and between 80% and 85% for the NPV system. Compared with the NPV system, the PNCV system increased the vertical airflow movement, causing significant cooling and dehumidifying effects. Hence, the proposed system provides an effective new ventilation mode for achieving efficient and accurate environmental control in laying hen cages. Full article
(This article belongs to the Section Animal System and Management)
Show Figures

Figure 1

Figure 1
<p>Ventilation structure: (<b>a</b>) schematic diagram of the PNCV system structure; (<b>b</b>) internal structure of the laying house; (<b>c</b>) exterior structure diagram of laying house.</p>
Full article ">Figure 2
<p>Negative-pressure-fan-structure schematic diagram.</p>
Full article ">Figure 3
<p>Evaporative cooling pad room: (<b>a</b>) structure diagram; (<b>b</b>) site structure diagram.</p>
Full article ">Figure 4
<p>Three-dimensional model of the laying hen house.</p>
Full article ">Figure 5
<p>Detailed view of the grid: (<b>a</b>) the distribution of the grid; (<b>b</b>) the measurement nodes of unrelated solutions.</p>
Full article ">Figure 6
<p>Temperature with different grid sizes.</p>
Full article ">Figure 7
<p>Control group experiment: schematic diagram of fan intake angle.</p>
Full article ">Figure 8
<p>Locations of Planes 1–4 used for the illustration of spatial variations of the indoor environment.</p>
Full article ">Figure 9
<p>On-site measurement points for environmental parameters.</p>
Full article ">Figure 10
<p>The difference between simulated results and the experimental measurements: (<b>a</b>) relative error of air velocity; (<b>b</b>) relative error of temperature; (<b>c</b>) relative error of relative humidity.</p>
Full article ">Figure 11
<p>Simulation results of air velocity: (<b>a</b>) average air velocity and standard deviation in the CZ; (<b>b</b>) proportion of the volume of different wind speeds in CZ.</p>
Full article ">Figure 12
<p>Air velocity distribution in the spatial region.</p>
Full article ">Figure 13
<p>Distribution of mass-weighted uniformity index (MWUI) of air velocity in the CZ.</p>
Full article ">Figure 14
<p>Temperature distribution at the center line of the house for different heights in: Y = 0.7 m, Y = 1.4 m, Y = 2.1 m, and Y = 2.8 m.</p>
Full article ">Figure 15
<p>Temperature distribution in the spatial region.</p>
Full article ">Figure 16
<p>Humidity distribution at the center line of the house for different heights in: Y = 0.7 m, Y = 1.4 m, Y = 2.1 m, and Y = 2.8 m.</p>
Full article ">Figure 17
<p>Relative humidity distribution in the spatial region.</p>
Full article ">
22 pages, 8904 KiB  
Article
A Y-Type Air-Cooled Battery Thermal Management System with a Short Airflow Path for Temperature Uniformity
by Xiangyang Li, Jing Liu and Xiaomin Li
Batteries 2024, 10(9), 302; https://doi.org/10.3390/batteries10090302 - 27 Aug 2024
Viewed by 507
Abstract
A Y-type air-cooled structure has been proposed to improve the heat dissipation efficiency and temperature uniformity of battery thermal management systems (BTMSs) by reducing the flow path of air. By combining computational fluid dynamics (CFD) methods, the influence of the depths of the [...] Read more.
A Y-type air-cooled structure has been proposed to improve the heat dissipation efficiency and temperature uniformity of battery thermal management systems (BTMSs) by reducing the flow path of air. By combining computational fluid dynamics (CFD) methods, the influence of the depths of the distribution and convergence plenums on the airflow velocity through battery cells was analyzed to improve heat dissipation efficiency. Adjusting the width of the first and ninth cooling channels can change the air velocity of these two channels, thereby improving the temperature uniformity of the BTMS. Further discussion was conducted regarding the influences of inlet and outlet depths. When the inlet width and outlet width were 20 mm, the maximum temperature and maximum temperature difference of the Y-type BTMS were 39.84 °C and 0.066 °C at a discharge rate of 2.5 °C, respectively; these temperatures were 1.537 °C (3.68%) and 0.059 °C (47.2%) lower than those of the T-type model. Meanwhile, the energy consumption of the sample also decreased by 13.1%. The results indicate that the heat dissipation performance of the proposed Y-type BTMS was improved, achieving excellent temperature uniformity, and the energy consumption was also reduced. Full article
(This article belongs to the Special Issue Recent Advances in the Thermal Safety of Lithium-Ion Batteries)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structural design of BTMSs. (<b>a</b>) Z-type; (<b>b</b>) U-type; (<b>c</b>) T-type.</p>
Full article ">Figure 2
<p>Three-dimensional model of Y-type air-cooled BTMS and battery cell.</p>
Full article ">Figure 3
<p>Half of the 3D mesh model of the battery pack cut in cross-section normal to the <span class="html-italic">X</span>-axis.</p>
Full article ">Figure 4
<p>Comparison between the simulation results of a T-type air-cooled BTMS battery pack in this paper and the results given in reference [<a href="#B14-batteries-10-00302" class="html-bibr">14</a>]. Copyright (2024), with permission from Elsevier.</p>
Full article ">Figure 5
<p>Right and top views of a Y-type air-cooled BTMS.</p>
Full article ">Figure 6
<p>The impact of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>c</mi> </msub> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> are equal and range from 2.5 mm to 20 mm.</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>max</mi> </mrow> </msub> </mrow> </semantics></math> of the system and <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mrow> <mi>max</mi> </mrow> </msub> </mrow> </semantics></math> at the cross-sectional position of the system in the <span class="html-italic">X</span>-axis.</p>
Full article ">Figure 8
<p>The velocity cloud map located in the <span class="html-italic">X</span>-axis cross-section, where <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> are equal. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.5 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 6 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 7 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 10 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 20 mm.</p>
Full article ">Figure 8 Cont.
<p>The velocity cloud map located in the <span class="html-italic">X</span>-axis cross-section, where <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> are equal. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.5 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 6 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 7 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 10 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 20 mm.</p>
Full article ">Figure 9
<p>The heat dissipation performance of Y-type BTMS when <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> are not equal; the sum of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> is set to 11 mm.</p>
Full article ">Figure 10
<p>The system’s heat dissipation performance when the <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> was kept at 8.0 mm.</p>
Full article ">Figure 11
<p>The air flow velocity in the <span class="html-italic">X</span>-axis cross-section, where <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> is 8.0 mm. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 1.5 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.0 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3.0 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3.5 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.0 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.5 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.0 mm.</p>
Full article ">Figure 11 Cont.
<p>The air flow velocity in the <span class="html-italic">X</span>-axis cross-section, where <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> is 8.0 mm. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 1.5 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.0 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3.0 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 3.5 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.0 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 4.5 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> = 5.0 mm.</p>
Full article ">Figure 12
<p>The effect of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>c</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> on <math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mrow> <mi>max</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mrow> <mi>max</mi> </mrow> </msub> </mrow> </semantics></math> at the cross-section in the <span class="html-italic">X</span>-axis. <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>d</mi> <mi>p</mi> </mrow> </msub> </mrow> </semantics></math> is 8.0 mm.</p>
Full article ">Figure 13
<p>The influence of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> on the temperature of the battery pack.</p>
Full article ">Figure 14
<p>The system flow velocity cloud map with changes in <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math>, with the cross-section taken from the middle section of the battery pack in the <span class="html-italic">Y</span>-axis direction. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 1.7 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 1.9 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.1 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.2 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.3 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.7 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 3.0 mm.</p>
Full article ">Figure 14 Cont.
<p>The system flow velocity cloud map with changes in <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math>, with the cross-section taken from the middle section of the battery pack in the <span class="html-italic">Y</span>-axis direction. (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 1.7 mm; (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 1.9 mm; (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.1 mm; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.2 mm; (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.3 mm; (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.5 mm; (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 2.7 mm; (<b>h</b>) <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> = 3.0 mm.</p>
Full article ">Figure 15
<p>The maximum flow velocities of cooling channel 1 and all cooling channels when <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> change, with <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>1</mn> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mn>9</mn> </msub> </mrow> </semantics></math> ranging from 1.7 mm to 3.0 mm.</p>
Full article ">Figure 16
<p>The impact of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics></math> on system performance.</p>
Full article ">Figure 17
<p>The effect of <math display="inline"><semantics> <mrow> <msub> <mi>d</mi> <mrow> <mi>o</mi> <mi>u</mi> <mi>t</mi> </mrow> </msub> </mrow> </semantics></math> on system performance.</p>
Full article ">Figure 18
<p><math display="inline"><semantics> <mrow> <msub> <mi>T</mi> <mi>c</mi> </msub> </mrow> </semantics></math> of battery cell 4 at different discharge rates. The ambient temperature is 25 °C.</p>
Full article ">
20 pages, 4364 KiB  
Article
Numerical Study of Melt-Spinning Dynamic Parameters and Microstructure Development with Ongoing Crystallization
by Xiangqian Liu, Pei Feng, Chongchang Yang and Zexu Hu
Polymers 2024, 16(17), 2398; https://doi.org/10.3390/polym16172398 - 23 Aug 2024
Viewed by 507
Abstract
In response to an investigation on the paths of changes in the crystallization and radial differences during the forming process of nascent fibers, in this study, we conducted numerical simulation and analyzed the changes in crystallization mechanical parameters and tensile properties through a [...] Read more.
In response to an investigation on the paths of changes in the crystallization and radial differences during the forming process of nascent fibers, in this study, we conducted numerical simulation and analyzed the changes in crystallization mechanical parameters and tensile properties through a fluid dynamics two-phase model. The model was based on the melt-spinning method focusing on melt spinning, the environment of POLYFLOW, and the method of joint simulation, coupled with Nakamura crystallization kinetics, including the development of process collaborative parameters, stretch-induced crystallization, viscoelasticity, filament cooling, gravity term, inertia, and air resistance. Finally, for nylon 6 BHS and CN9987 resin spinning, the model successfully predicted the distribution changes in temperature, velocity, strain rate tensor, birefringence, and stress tensor along the axial and radial fibers and obtained the variation pattern of fibers’ crystallinity along the entire spinning process under different stretching rates. Furthermore, we also explored the effects of spinning conditions, including inlet flow rate, winding speeds, and the extrusion temperature, on the fibers’ crystallization process and obtained the influence rules of different spinning conditions on fiber crystallization. Knowing the paths of changes in mechanical performance can provide important guidance and optimization strategies for the future industrial preparation of high-performance fibers. Full article
(This article belongs to the Section Polymer Physics and Theory)
Show Figures

Figure 1

Figure 1
<p>Process of melt spinning fiber (the computational domain covers the entire enclosed area from the fluid to the inlet of the spinneret. <math display="inline"> <semantics> <mrow> <msub> <mo>Γ</mo> <mrow> <mi>i</mi> <mi>n</mi> </mrow> </msub> </mrow> </semantics> </math> is the inlet boundary of the fluid, <math display="inline"> <semantics> <mrow> <msub> <mo>Γ</mo> <mrow> <mi>wall</mi> </mrow> </msub> </mrow> </semantics> </math> is the wall surface of the spinneret, <math display="inline"> <semantics> <mrow> <msub> <mo>Γ</mo> <mrow> <mi>out</mi> </mrow> </msub> </mrow> </semantics> </math> is the outlet of the fluid, <math display="inline"> <semantics> <mrow> <msub> <mo>Γ</mo> <mrow> <mi>free</mi> </mrow> </msub> </mrow> </semantics> </math> is the free surface of the fiber).</p>
Full article ">Figure 2
<p>Block diagram of model solution.</p>
Full article ">Figure 3
<p>Model and grid division.</p>
Full article ">Figure 4
<p>Temperature variation of (<b>a</b>) different constitutive equations; (<b>b</b>) different take-up speeds throughout the spinning process [<a href="#B35-polymers-16-02398" class="html-bibr">35</a>,<a href="#B36-polymers-16-02398" class="html-bibr">36</a>].</p>
Full article ">Figure 5
<p>Speed variation throughout the spinning process. (<b>a</b>) Contrast of different models [<a href="#B15-polymers-16-02398" class="html-bibr">15</a>]; (<b>b</b>) contrast of simulation and experiment [<a href="#B37-polymers-16-02398" class="html-bibr">37</a>] (<span class="html-italic">Q</span> = 3 × 10<sup>−8</sup> m<sup>3</sup>/s, <span class="html-italic">T</span> = 533 k, <span class="html-italic">T</span><sub>0</sub> = 293 k, <span class="html-italic">V</span><sub>1</sub> = 10 m/s, <span class="html-italic">V</span><sub>2</sub> = 30 m/s, <span class="html-italic">V</span><sub>0</sub> = 0.4 m/s, <span class="html-italic">L</span> = 2 m).</p>
Full article ">Figure 6
<p>Comparison of velocity distribution at speeds of 30 m/s–70 m/s. (<b>a</b>) Axial distribution of take-up velocity along the spinning process; (<b>b</b>) radial distribution of velocity at different z points (<span class="html-italic">Q</span> = 4.5 × 10<sup>−8</sup> m<sup>3</sup>/s, <span class="html-italic">T</span> = 533 k, <span class="html-italic">T</span><sub>0</sub> = 293 k, <span class="html-italic">V</span><sub>0</sub> = 0.4 m/s, <span class="html-italic">L</span> = 2 m).</p>
Full article ">Figure 7
<p>Distribution of radial temperature at different z positions of fibers. (<b>a</b>) Initial deformation zone (<span class="html-italic">z</span> value 5 mm–150 mm); (<b>b</b>) solidification zone (<span class="html-italic">z</span> value 1300 mm–1600 mm).</p>
Full article ">Figure 8
<p>Prediction of fiber surface with different take-up velocities. (<b>a</b>) Comparison of the predicted free surface and experimental measurement [<a href="#B13-polymers-16-02398" class="html-bibr">13</a>,<a href="#B37-polymers-16-02398" class="html-bibr">37</a>]; (<b>b</b>) changes in the surface of the filament in the swollen area (<span class="html-italic">Q</span> = 4.5 × 10<sup>−8</sup> m<sup>3</sup>/s, <span class="html-italic">T</span> = 533 k, <span class="html-italic">T</span><sub>0</sub> = 293 k, <span class="html-italic">V</span><sub>0</sub> = 0.4 m/s, <span class="html-italic">L</span> = 2 m).</p>
Full article ">Figure 9
<p>Effect of different take-up velocities on fiber tensor distribution. (<b>a</b>) Average stress; (<b>b</b>) radial distribution of total external stress tensor at different z points; (<b>c</b>) distribution of stress component in the stretching z direction; (<b>d</b>) distribution of strain rate tensor along the spinning process.</p>
Full article ">Figure 9 Cont.
<p>Effect of different take-up velocities on fiber tensor distribution. (<b>a</b>) Average stress; (<b>b</b>) radial distribution of total external stress tensor at different z points; (<b>c</b>) distribution of stress component in the stretching z direction; (<b>d</b>) distribution of strain rate tensor along the spinning process.</p>
Full article ">Figure 10
<p>Birefringence distribution of silk slips at different drafts throughout the spinning process [<a href="#B38-polymers-16-02398" class="html-bibr">38</a>] (<span class="html-italic">V</span><sub>d1</sub> = 30 m/s, <span class="html-italic">V</span><sub>d2</sub> = 40 m/s, <span class="html-italic">V</span><sub>d3</sub> = 50 m/s, <span class="html-italic">V</span><sub>d4</sub> = 60 m/s, <span class="html-italic">V</span><sub>d5</sub> = 70 m/s).</p>
Full article ">Figure 11
<p>Effect of different spinning conditions on fiber crystallinity. (<b>a</b>) Different take-up velocities; (<b>b</b>) relationship between take-up velocities and crystallinity; (<b>c</b>) inlet flow rate; (<b>d</b>) spinneret wall temperature.</p>
Full article ">Figure 11 Cont.
<p>Effect of different spinning conditions on fiber crystallinity. (<b>a</b>) Different take-up velocities; (<b>b</b>) relationship between take-up velocities and crystallinity; (<b>c</b>) inlet flow rate; (<b>d</b>) spinneret wall temperature.</p>
Full article ">Figure 12
<p>Distribution of radial temperature <span class="html-italic">T</span>, stress tensor <math display="inline"> <semantics> <mi>τ</mi> </semantics> </math>, crystallinity <math display="inline"> <semantics> <mi>θ</mi> </semantics> </math>, and birefringence <math display="inline"> <semantics> <mrow> <mo>Δ</mo> <mi>n</mi> </mrow> </semantics> </math> (<math display="inline"> <semantics> <mrow> <msub> <mi>X</mi> <mrow> <mi>max</mi> </mrow> </msub> </mrow> </semantics> </math> represents the maximum value of <span class="html-italic">X</span>).</p>
Full article ">
13 pages, 6168 KiB  
Article
Cooling Air Velocity on Iron Ore Pellet Performance Based on Experiments and Simulations
by Liming Ma, Jianliang Zhang, Zhengjian Liu, Qiuye Cai, Liangyuan Hao, Shaofeng Lu, Huiqing Jiang and Yaozu Wang
Metals 2024, 14(8), 919; https://doi.org/10.3390/met14080919 - 14 Aug 2024
Viewed by 546
Abstract
During the pellet cooling process, cooling air velocity is crucial for optimizing the cooling rate, evaluating the utilization rate of cooling heat energy, and improving pellet performance. As the simulated cooling air velocity increased, the gas temperature at the cooling endpoint decreased from [...] Read more.
During the pellet cooling process, cooling air velocity is crucial for optimizing the cooling rate, evaluating the utilization rate of cooling heat energy, and improving pellet performance. As the simulated cooling air velocity increased, the gas temperature at the cooling endpoint decreased from 87 °C to 51 °C, and the solid temperature decreased from 149 °C to 103 °C. The total enthalpy of the recovered gas initially reduced and then increased while the heat recovery rate gradually increased. During the experiment, the inhomogeneity of pellet quality gradually increased with the rise in cooling air velocity. The effect of cooling air velocity on pellet properties is primarily reflected in the formation of cracks and low-melting liquid phases (FeO and fayalite). As the cooling air velocity increases, the softening onset temperature of the pellet decreases significantly. The melting zone decreases from 193 °C to 105 °C, and the permeability of the adhesive zone increases. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the cooling experiment for the iron ore pellets ((<b>a</b>) green pellet preparation; (<b>b</b>) drying of green pellets; (<b>c</b>) preheating and roasting; (<b>d</b>) cooling of iron ore pellets).</p>
Full article ">Figure 2
<p>Schematic diagram of cooling simulation of the iron ore pellets ((<b>a</b>) schematic diagram of pellet cooling equipment; (<b>b</b>) modelling of iron ore pellet cooling; (<b>c</b>) iron ore pellet cooling baseline model; (<b>d</b>) accuracy calibration of cooling models of iron ore pellet).</p>
Full article ">Figure 3
<p>Influence of cooling air velocity on pellet performance ((<b>a</b>) cooling time; (<b>b</b>) pellet strength).</p>
Full article ">Figure 4
<p>Light microscopic analysis of pellets at different cooling air velocities ((<b>a</b>–<b>d</b>) 2 m<sup>3</sup>/h; (<b>e</b>–<b>h</b>) 4 m<sup>3</sup>/h; (<b>i</b>–<b>l</b>) 6 m<sup>3</sup>/h).</p>
Full article ">Figure 5
<p>SEM-EDS analysis positions of pellets at different cooling air velocities ((<b>a</b>–<b>c</b>) 2 m<sup>3</sup>/h; (<b>d</b>–<b>f</b>) 4 m<sup>3</sup>/h; (<b>g</b>–<b>i</b>) 6 m<sup>3</sup>/h).</p>
Full article ">Figure 6
<p>Effect of cooling gas on gas–solid phase temperature ((<b>a</b>) gas temperature; (<b>b</b>) temperature of pellet).</p>
Full article ">Figure 7
<p>Effect of cooling air velocity on cooling rate and enthalpy ((<b>a</b>) cooling rate; (<b>b</b>) enthalpy).</p>
Full article ">Figure 8
<p>Effect of cooling air velocity on heat recovery.</p>
Full article ">Figure 9
<p>Softening–melting characteristics of pellets at different cooling air velocities.</p>
Full article ">
23 pages, 9533 KiB  
Article
Experimental Investigation on the Damage Evolution of Thermally Treated Granodiorite Subjected to Rapid Cooling with Liquid Nitrogen
by Mohamed Elgharib Gomah, Enyuan Wang and Ahmed A. Omar
Sustainability 2024, 16(15), 6396; https://doi.org/10.3390/su16156396 - 26 Jul 2024
Viewed by 654
Abstract
In many thermal geotechnical applications, liquid nitrogen (LN2) utilization leads to damage and cracks in the host rock. This phenomenon and associated microcracking are a hot topic that must be thoroughly researched. A series of physical and mechanical experiments were conducted [...] Read more.
In many thermal geotechnical applications, liquid nitrogen (LN2) utilization leads to damage and cracks in the host rock. This phenomenon and associated microcracking are a hot topic that must be thoroughly researched. A series of physical and mechanical experiments were conducted on Egyptian granodiorite samples to investigate the effects of liquid nitrogen cooling on the preheated rock. Before quenching in LN2, the granodiorite was gradually heated to 600 °C for two hours. Microscopical evolution was linked to macroscopic properties like porosity, mass, volume, density, P-wave velocity, uniaxial compressive strength, and elastic modulus. According to the experiment results, the thermal damage, crack density, porosity, and density reduction ratio increased gradually to 300 °C before severely degrading beyond this temperature. The uniaxial compressive strength declined marginally to 200 °C, then increased to 300 °C before monotonically decreasing as the temperature rose. On the other hand, at 200 °C, the elastic modulus and P-wave velocity started to decline significantly. Thus, 200 and 300 °C were noted in this study as two mutation temperatures in the evolution of granodiorite mechanical and physical properties, after which all parameters deteriorated. Moreover, LN2 cooling causes more remarkable physical and mechanical modifications at the same target temperature than air cooling. Through a deeper comprehension of how rocks behave in high-temperature conditions, this research seeks to avoid and limit future geological risks while promoting sustainability and understanding the processes underlying rock failure. Full article
Show Figures

Figure 1

Figure 1
<p>An Eastern Desert geological map of Egypt that includes the research area [<a href="#B53-sustainability-16-06396" class="html-bibr">53</a>].</p>
Full article ">Figure 2
<p>The XRD patterns of granodiorite samples at ambient temperature and following different treatments.</p>
Full article ">Figure 3
<p>The principal devices used in this study: (<b>a</b>) the WiseTherm electric furnace, (<b>b</b>) quenched samples of liquid nitrogen, (<b>c</b>) the SEM apparatus, (<b>d</b>) the optical microscope device, and (<b>e</b>) a uniaxial compression test unit.</p>
Full article ">Figure 4
<p>Porosity and water absorption responses of granodiorite samples following various thermal treatments and LN<sub>2</sub> cooling.</p>
Full article ">Figure 5
<p>Mass loss, volume increase, and density reduction rates of granodiorite with target temperatures.</p>
Full article ">Figure 6
<p>Change in P-wave velocity of granodiorite samples with temperature with LN<sub>2</sub> cooling treatments.</p>
Full article ">Figure 7
<p>SEM images of granodiorite thermally treated and quickly cooled via LN<sub>2</sub> at 200 °C (<b>a</b>), 300 °C (<b>b</b>), 400 °C (<b>c</b>), and 600 °C (<b>d</b>). (bc) is boundary microcracks, and (tc) is transgranular microcracks.</p>
Full article ">Figure 8
<p>Crack length and width evolution of granodiorite specimens subjected to various thermal treatments and LN<sub>2</sub> rapid cooling.</p>
Full article ">Figure 9
<p>Diagram of the microstructural changes in studied samples after heat treatment, followed by rapid cooling using LN<sub>2</sub>.</p>
Full article ">Figure 10
<p>The change in the uniaxial compressive strength (UCS) of granodiorite after heat treatment and LN<sub>2</sub> cooling.</p>
Full article ">Figure 11
<p>The elastic modulus (E) of granodiorite as a function of temperature and LN<sub>2</sub> cooling approach.</p>
Full article ">Figure 12
<p>Physical and mechanical property responses during heating and LN<sub>2</sub> cooling processes.</p>
Full article ">Figure 13
<p>Crack evolution in granodiorite specimens subjected to different thermal treatments (<b>a</b>) at 200 °C and (<b>b</b>) at 300 °C and rapid cooling with LN<sub>2</sub>, as seen using an optical microscope.</p>
Full article ">Figure 14
<p>Crack evolution in granodiorite specimens subjected to different thermal treatments (<b>a</b>) at 400 °C and (<b>b</b>) at 600 °C and rapid cooling with LN<sub>2</sub>, as seen using an optical microscope.</p>
Full article ">Figure 15
<p>Relation between P-wave velocity and porosity of thermally heated granodiorite, followed by the LN<sub>2</sub> cooling method.</p>
Full article ">Figure 16
<p>Thermal damage evolution of P-wave velocity and elastic modulus of thermally heated granodiorite, followed by the LN<sub>2</sub> cooling approach.</p>
Full article ">Figure 17
<p>Cooling impact analysis on the physical properties of thermally heated granodiorite at various temperatures is followed by air cooling (A-C) and liquid nitrogen (LN<sub>2</sub>).</p>
Full article ">Figure 18
<p>Cooling impact analysis on the mechanical properties of thermally heated granodiorite at various temperatures is followed by air cooling (A-C) and liquid nitrogen cooling (LN<sub>2</sub>).</p>
Full article ">
16 pages, 4023 KiB  
Article
Experimental Study on Heat Transfer Characteristics of Radiant Cooling and Heating
by Shengpeng Chen, Xiaohui Ma and Chaoling Han
Energies 2024, 17(13), 3304; https://doi.org/10.3390/en17133304 - 5 Jul 2024
Viewed by 574
Abstract
While traditional air conditioning systems serve their purpose, radiation air conditioning systems provide several benefits, including improved comfort, higher energy efficiency, and lower initial costs. Nevertheless, the heat exchange capacity per unit area of the radiation plate in such systems is somewhat restricted, [...] Read more.
While traditional air conditioning systems serve their purpose, radiation air conditioning systems provide several benefits, including improved comfort, higher energy efficiency, and lower initial costs. Nevertheless, the heat exchange capacity per unit area of the radiation plate in such systems is somewhat restricted, which directly affects their practical engineering applications. To address this, experimental investigations were undertaken to examine the impact of cold/hot water supply temperature, water flow velocity, and surface emissivity of radiant panels on their heat transfer characteristics for both summer cooling and winter heating. The findings highlight the significant influence of water supply temperature, flow rate, and surface emissivity on the heat transfer properties of the radiant plates. It is worth noting that adjustments to the water flow rate and surface emissivity impose limitations on enhancing the radiant plate heat transfer performance. For instance, in summer, the heat transfer coefficient of the roughly machined light alumina plate radiant panel was determined by fitting the experimental heat transfer data against characteristic temperatures. Specifically, during cooling, the total heat transfer coefficient of the radiant plate was calculated as 6.77 W/(m2·K), comprising a thermal coefficient of 5.41 W/(m2·K) and a convective heat transfer coefficient of 4.17 W/(m2·K). Conversely, during winter heating, the total heat transfer coefficient of the radiant plate increased to 8.94 W/(m2·K), with a radiation heat transfer coefficient of 6.13 W/(m2·K) and a convective heat transfer coefficient of 3.79 W/(m2·K). Full article
(This article belongs to the Section J1: Heat and Mass Transfer)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the radiant plate structure. 1—surface layer; 2—insulation layer; 3—main pipe; 4—branch pipe.</p>
Full article ">Figure 2
<p>Appearance of the radiant panel. (<b>a</b>) roughly processed slight alumina plate, (<b>b</b>) black spray-painted surface aluminum plate, (<b>c</b>) white spray-painted surface aluminum plate.</p>
Full article ">Figure 3
<p>Overview of the experimental monitoring system.</p>
Full article ">Figure 4
<p>Influence of different cold water supply temperatures on heat transfer characteristics of the radiant plate [<a href="#B25-energies-17-03304" class="html-bibr">25</a>].</p>
Full article ">Figure 5
<p>Influence of different flow rates of cold water supply on the heat transfer characteristics of the radiant panels.</p>
Full article ">Figure 6
<p>Influence of different surface emissivity values on heat transfer characteristics of the radiant plate in summer.</p>
Full article ">Figure 7
<p>Diagram of the radiant plate heat transfer characteristics in summer (Arrow in the figure refer to the temperature range).</p>
Full article ">Figure 7 Cont.
<p>Diagram of the radiant plate heat transfer characteristics in summer (Arrow in the figure refer to the temperature range).</p>
Full article ">Figure 8
<p>Influence of different hot water supply temperatures on heat transfer characteristics of the radiant plate.</p>
Full article ">Figure 9
<p>Influence of different hot water supply flow rates on the heat transfer characteristics of the radiant panels.</p>
Full article ">Figure 10
<p>Influence of different surface emissivity on heat transfer characteristics of the radiant plate in winter.</p>
Full article ">Figure 11
<p>Diagram of the radiant plate heat transfer characteristics in winter.</p>
Full article ">
14 pages, 4985 KiB  
Article
Bénard–Marangoni Convection in an Open Cavity with Liquids at Low Prandtl Numbers
by Hao Jiang, Wang Liao and Enhui Chen
Symmetry 2024, 16(7), 844; https://doi.org/10.3390/sym16070844 - 4 Jul 2024
Viewed by 586
Abstract
Bénard–Marangoni convection in an open cavity has attracted much attention in the past century. In most of the previous works, liquids with Prandtl numbers larger than unity were used to study in this issue. However, the Bénard–Marangoni convection with liquids at Prandtl numbers [...] Read more.
Bénard–Marangoni convection in an open cavity has attracted much attention in the past century. In most of the previous works, liquids with Prandtl numbers larger than unity were used to study in this issue. However, the Bénard–Marangoni convection with liquids at Prandtl numbers lower than unity is still unclear. In this study, Bénard–Marangoni convection in an open cavity with liquids at Prandtl numbers lower than unity in zero-gravity conditions is investigated to reveal the bifurcations of the flow and quantify the heat and mass transfer. Three-dimensional direct numerical simulation is conducted by the finite-volume method with a SIMPLE scheme for the pressure–velocity coupling. The bottom boundary is nonslip and isothermal heated. The top boundary is assumed to be flat, cooled by air and opposed by the Marangoni stress. Numerical simulation is conducted for a wide range of Marangoni numbers (Ma) from 5.0 × 101 to 4.0 × 104 and different Prandtl numbers (Pr) of 0.011, 0.029, and 0.063. Generally, for small Ma, the liquid metal in the cavity is dominated by conduction, and there is no convection. The critical Marangoni number for liquids with Prandtl numbers lower than unity equals those with Prandtl numbers larger than unity, but the cells are different. As Ma increases further, the cells pattern becomes irregular and the structure of the top surface of the cells becomes finer. The thermal boundary layer becomes thinner, and the column of velocity magnitudes in the middle slice of the fluid is denser, indicating a stronger convection with higher Marangoni numbers. A new scaling is found for the area-weighted mean velocity magnitude at the top boundary of um~Ma Pr−2/3, which means the mass transfer may be enhanced by high Marangoni numbers and low Prandtl numbers. The Nusselt number is approximately constant for Ma ≤ 400 but increases slowly for Ma > 400, indicating that the heat transfer may be enhanced by increasing the Marangoni number. Full article
(This article belongs to the Special Issue Symmetry and Its Applications in Experimental Fluid Mechanics)
Show Figures

Figure 1

Figure 1
<p>Schematic of physical model.</p>
Full article ">Figure 2
<p>The temperature and velocity magnitude for cases with <span class="html-italic">Ma</span> &lt;&lt; <span class="html-italic">Ma<sub>c</sub></span>. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 3
<p>The temperature and velocity magnitude for cases with <span class="html-italic">Ma</span> = 80 and <span class="html-italic">Pr</span> = 0.029. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 4
<p>The temperature and velocity magnitude for cases with <span class="html-italic">ε</span> = 0.25 and <span class="html-italic">Pr</span> = 0.029. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 5
<p>The temperature and velocity magnitude for cases with <span class="html-italic">ε</span> = 1.5 and <span class="html-italic">Pr</span> = 0.029. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 6
<p>The temperature and velocity magnitude for cases with <span class="html-italic">ε</span> = 36.5 and <span class="html-italic">Pr</span> = 0.029. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 7
<p>The temperature and velocity magnitude for cases with <span class="html-italic">ε</span> = 499 and <span class="html-italic">Pr</span> = 0.029. (<b>a</b>) The temperature field on the top surface. (<b>b</b>) The velocity magnitude on the top surface. (<b>c</b>) The temperature field at the slice of <span class="html-italic">x</span> = 0. (<b>d</b>) The velocity magnitude at the slice of <span class="html-italic">x</span> = 0. Note that the height of slice is stretched for better visualization.</p>
Full article ">Figure 8
<p>The area-weighted average velocity magnitude of the top boundary.</p>
Full article ">Figure 9
<p>The Nusselt number on the heated bottom boundary.</p>
Full article ">
29 pages, 22049 KiB  
Article
Predicting Erosion Damage in a Centrifugal Fan
by Adel Ghenaiet
Int. J. Turbomach. Propuls. Power 2024, 9(2), 23; https://doi.org/10.3390/ijtpp9020023 - 17 Jun 2024
Viewed by 863
Abstract
Erosion damage can occur in fans and blowers during industrial processes, cooling, and mine ventilation. This study focuses on investigating erosion caused by particulate air flows in a centrifugal fan with forward-inclined blades. This type of fan is particularly vulnerable to erosion due [...] Read more.
Erosion damage can occur in fans and blowers during industrial processes, cooling, and mine ventilation. This study focuses on investigating erosion caused by particulate air flows in a centrifugal fan with forward-inclined blades. This type of fan is particularly vulnerable to erosion due to its radial flow component and flow recirculation. The flow field was solved separately, and the data transferred to the particle trajectory and erosion code. This in-house code implements the Lagrangian approach and the random walk algorithm, including statistical descriptions of particle sizes, release positions, and restitution factors. The study involved two types of dust particles, with a concentration between 100 and 500 μg/m3: The first type is the Saharan (North Africa) dust, which has a finer size between 0.1 and 100 microns. The second type is the Coarse Arizona Road Dust, also known as AC-coarse dust, which has a larger size ranging from 1 to 200 microns. The complex flow conditions within the impeller and scroll, as well as the concentration and size distribution of particles, are shown to affect the paths, impact conditions, and erosion patterns. The outer wall of the scroll is most heavily eroded due to high-impact velocities by particles exiting the impeller. Erosion is more pronounced on the pressure side of the full blades compared to the splitters and casing plate. The large non-uniformities of erosion patterns indicate a strong dependence with the blade position around the scroll. Therefore, the computed eroded mass is cumulated and averaged for all the surfaces of components. These results provide useful insights for monitoring erosion wear in centrifugal fans and selecting appropriate coatings to extend the lifespan. Full article
Show Figures

Figure 1

Figure 1
<p>Centrifugal fan (<b>left</b>) and impeller (<b>right</b>).</p>
Full article ">Figure 2
<p>Computational domain.</p>
Full article ">Figure 3
<p>Meshing of the (<b>a</b>) impeller and (<b>b</b>) scroll tongue region.</p>
Full article ">Figure 4
<p>Distribution of <span class="html-italic">y</span><sup>+</sup>: (<b>a</b>) impeller and (<b>b</b>) scroll.</p>
Full article ">Figure 5
<p>Grid size independence verification.</p>
Full article ">Figure 6
<p>Computed performance compared with reference (dashed line).</p>
Full article ">Figure 7
<p>Static pressure at (<b>a</b>) mid-span and (<b>b</b>) near blade tip.</p>
Full article ">Figure 8
<p>Flow velocity at the meridional plane halving the fan components.</p>
Full article ">Figure 9
<p>Flow velocity at (<b>a</b>) mid-span and (<b>b</b>) near blade tip.</p>
Full article ">Figure 10
<p>Meridional flow velocities near (<b>a</b>) the exit from the scroll and (<b>b</b>) tongue.</p>
Full article ">Figure 11
<p>Flow structures at cross-sections of the impeller.</p>
Full article ">Figure 12
<p>Flow streamlines.</p>
Full article ">Figure 13
<p>Turbulent kinetic energy.</p>
Full article ">Figure 14
<p>Impact conditions.</p>
Full article ">Figure 15
<p>Dust particle size distributions.</p>
Full article ">Figure 16
<p>Release positions and sizes of Saharan dust particles at the highest concentration for two randomness factors.</p>
Full article ">Figure 17
<p>Samples of particle (1 μm) trajectories coloured by velocity: (<b>a</b>) top view and (<b>b</b>) side view.</p>
Full article ">Figure 18
<p>Samples of particle (100 μm) trajectories coloured by velocity: (<b>a</b>) top view and (<b>b</b>) side view.</p>
Full article ">Figure 19
<p>Trajectories of Saharan dust (0.1–100 microns) colored by (<b>a</b>) the particle diameter, (<b>b</b>) side view, and (<b>c</b>) Stokes number.</p>
Full article ">Figure 20
<p>Erosion with the blades’ positions, caused by Saharan dust (0.1–100 microns) of the concentration 500 μg/m<sup>3</sup>: impeller blades, casing plate, and volute.</p>
Full article ">Figure 21
<p>EMPH with blades’ positions caused by Saharan dust (0.1–100 microns) of the concentration 500 μg/m<sup>3</sup>: (<b>a</b>) full blades, (<b>b</b>) splitters, (<b>c</b>) impeller, (<b>d</b>) casing plate, and (<b>e</b>) scroll.</p>
Full article ">Figure 22
<p>Erosion patterns in the impeller, caused by Saharan dust (0.1–100 microns) of the concentration 500 μg/m<sup>3</sup>, operating at (<b>a</b>) the nominal flow rate and (<b>b</b>) maximum discharge.</p>
Full article ">Figure 23
<p>Erosion patterns in the impeller, caused by AC-coarse dust (1–200 microns) of the concentration 500 μg/m<sup>3</sup>, operating at (<b>a</b>) the nominal flow rate and (<b>b</b>) maximum discharge.</p>
Full article ">Figure 24
<p>Erosion rate density in the scroll, caused by Saharan (0.1–100 microns) of the concentration 500 μg/m<sup>3</sup>, operating at (<b>a</b>) the nominal flow rate and (<b>b</b>) maximum discharge.</p>
Full article ">Figure 25
<p>Erosion rate density (mg/s·mm<sup>2</sup>) in the scroll, caused by AC-coarse dust (1–200 microns) of the concentration 500 μg/m<sup>3</sup>, operating at (<b>a</b>) the nominal flow rate and (<b>b</b>) maximum discharge.</p>
Full article ">Figure 25 Cont.
<p>Erosion rate density (mg/s·mm<sup>2</sup>) in the scroll, caused by AC-coarse dust (1–200 microns) of the concentration 500 μg/m<sup>3</sup>, operating at (<b>a</b>) the nominal flow rate and (<b>b</b>) maximum discharge.</p>
Full article ">Figure 26
<p>EMPH (mg/h) of the impeller, caused by (<b>a</b>) Saharan dust (0.1–100 microns) and (<b>b</b>) AC-coarse dust (1–200 microns).</p>
Full article ">Figure 27
<p>EMPH (mg/h) of centrifugal fan parts, caused by (<b>a</b>) Saharan dust (0.1–100 microns) and (<b>b</b>) AC-coarse dust (1–200 microns).</p>
Full article ">
20 pages, 6979 KiB  
Article
Multi-Strategical Thermal Management Approach for Lithium-Ion Batteries: Combining Forced Convection, Mist Cooling, Air Flow Improvisers and Additives
by Anikrishnan Mohanan and Kannan Chidambaram
World Electr. Veh. J. 2024, 15(5), 213; https://doi.org/10.3390/wevj15050213 - 11 May 2024
Cited by 1 | Viewed by 950
Abstract
Maintaining the peak temperature of a battery within limits is a mandate for the safer operation of electric vehicles. In two-wheeler electric vehicles, the options available for the battery thermal management system are minuscule due to the restrictions imposed by factors like weight, [...] Read more.
Maintaining the peak temperature of a battery within limits is a mandate for the safer operation of electric vehicles. In two-wheeler electric vehicles, the options available for the battery thermal management system are minuscule due to the restrictions imposed by factors like weight, cost, availability, performance, and load. In this study, a multi-strategical cooling approach of forced convection and mist cooling over a single-cell 21,700 lithium-ion battery working under the condition of 4C is proposed. The chosen levels for air velocities (10, 15, 20 and 25 m/s) imitate real-world riding conditions, and for mist cooling implementation, injection pressure with three levels (3, 7 and 14 bar) is considered. The ANSYS fluent simulation is carried out using the volume of fluid in the discrete phase modelling transition using water mist as a working fluid. Initial breakup is considered for more accurate calculations. The battery’s state of health (SOH) is determined using PYTHON by adopting the Newton–Raphson estimation. The maximum temperature reduction potential by employing an airflow improviser (AFI) and additives (Tween 80, 1-heptanol, APG0810, Tween 20 and FS3100) is also explored. The simulation results revealed that an additional reduction of about 11% was possible by incorporating additives and AFI in the multi-strategical approach. The corresponding SOH improvement was about 2%. When the electric two-wheeler operated under 4C, the optimal condition (Max. SOH and Min. peak cell temp.) was achieved at an air velocity of 25 m/s, injection pressure of 7 bar with AFI and 3% (by wt.) Tween 80 and a 0.1% deformer. Full article
(This article belongs to the Special Issue Thermal Management System for Battery Electric Vehicle)
Show Figures

Figure 1

Figure 1
<p>Design specifications, boundary conditions and assumptions for simulation.</p>
Full article ">Figure 2
<p>Nozzle specifications and its spray pattern.</p>
Full article ">Figure 3
<p>Impact of air velocities on maximum cell temperature and SOH.</p>
Full article ">Figure 4
<p>Impact of mist cooling injection pressure on maximum cell temperature and SOH.</p>
Full article ">Figure 5
<p>Contour plot of temperature for the battery cell (V = 25 m/s, P = 7 bar).</p>
Full article ">Figure 6
<p>Influence of forced air velocity over peak temperature.</p>
Full article ">Figure 7
<p>Reduced influence of mist injection at the bottom surface cell due to larger air velocity.</p>
Full article ">Figure 8
<p>Flow pattern where injection pressure overwhelms air inlet velocity.</p>
Full article ">Figure 9
<p>Temperature contour at a high injection pressure of 14 bar.</p>
Full article ">Figure 10
<p>Influence of injection pressure over different air velocities.</p>
Full article ">Figure 11
<p>Correlational chat of cell peak temperature and SOH.</p>
Full article ">Figure 12
<p>Reconstructed control volume with AFI (<b>a</b>) isometric view and (<b>b</b>) top view.</p>
Full article ">Figure 13
<p>(<b>a</b>) Re-designed control volume with AFI. (<b>b</b>) Flow path lines of particles showing increased turbulence and (<b>c</b>) velocity contour after 1 s.</p>
Full article ">Figure 14
<p>Simulated temperature contours in a combinational approach when (<b>a</b>) V = 15 m/s, P = 7 bar, (<b>b</b>) V = 20 m/s, P = 7 bar, (<b>c</b>) V = 25 m/s; P = 3.5 bar, (<b>d</b>) V = 25 m/s; P = 7 bar, and (<b>e</b>) V = 25 m/s; P = 7 bar with AFI.</p>
Full article ">Figure 15
<p>Impact of AFI on the cell peak temperature over a range of air velocities and injection pressures.</p>
Full article ">Figure 16
<p>Cell peak temperature and SOH variations due to AFI.</p>
Full article ">Figure 17
<p>Impact of additive addition on cell temperature for (<b>a</b>) the case of maximum reduction and (<b>b</b>) the case of minimum reduction (<b>c</b>,<b>d</b>) with the path line contour showing the increased turbulence distribution.</p>
Full article ">Figure 18
<p>Influence of additive type on peak cell temperature and SOH.</p>
Full article ">
Back to TopTop