[go: up one dir, main page]

Next Issue
Volume 4, September
Previous Issue
Volume 4, March
 
 

Dynamics, Volume 4, Issue 2 (June 2024) – 14 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Select all
Export citation of selected articles as:
24 pages, 4841 KiB  
Article
A Comparative Study of Different CFD Codes for Fluidized Beds
by Parindra Kusriantoko, Per Fredrik Daun and Kristian Etienne Einarsrud
Dynamics 2024, 4(2), 475-498; https://doi.org/10.3390/dynamics4020025 - 16 Jun 2024
Viewed by 1181
Abstract
Fluidized beds are pivotal in the process industry and chemical engineering, with Computational Fluid Dynamics (CFD) playing a crucial role in their design and optimization. Challenges in CFD modeling stem from the scarcity or inconsistency of experimental data for validation, along with the [...] Read more.
Fluidized beds are pivotal in the process industry and chemical engineering, with Computational Fluid Dynamics (CFD) playing a crucial role in their design and optimization. Challenges in CFD modeling stem from the scarcity or inconsistency of experimental data for validation, along with the uncertainties introduced by numerous parameters and assumptions across different CFD codes. This study navigates these complexities by comparing simulation results from the open-source MFIX and OpenFOAM, and the commercial ANSYS FLUENT, against experimental data. Utilizing a Eulerian–Eulerian framework and the kinetic theory of granular flow (KTGF), the investigation focuses on solid-phase properties through the classical drag laws of Gidaspow and Syamlal–O’Brien across varied parameters. Findings indicate that ANSYS Fluent, MFiX, and OpenFOAM can achieve reasonable agreement with experimental benchmarks, each showcasing distinct strengths and weaknesses. The study also emphasizes that both the Syamlal–O’Brien and Gidaspow drag models exhibit reasonable agreement with experimental benchmarks across the examined CFD codes, suggesting a moderated sensitivity to the choice of drag model. Moreover, analyses were also carried out for 2D and 3D simulations, revealing that the dimensional approach impacts the predictive accuracy to a certain extent, with both models adapting well to the complexities of each simulation environment. The study highlights the significant influence of restitution coefficients on bed expansion due to their effect on particle–particle collisions, with a value of 0.9 deemed optimal for balancing simulation accuracy and computational efficiency. Conversely, the specularity coefficient, impacting particle–wall interactions, exhibits a more subtle effect on bed dynamics. This finding emphasizes the critical role of carefully choosing these coefficients to effectively simulate the nuanced behaviors of fluidized beds. Full article
Show Figures

Figure 1

Figure 1
<p>The geometric setup used for simulations, adapted from Taghipour et al. [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Bed expansion for different superficial gas velocities with the Syamlal–O’Brien drag model in ANSYS Fluent, MFiX, and OpenFOAM. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 3
<p>Bed expansion for different superficial gas velocities with the Gidaspow drag model in ANSYS Fluent, MFiX, and OpenFOAM. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 4
<p>Pressure drop for different superficial gas velocities with the Syamlal–O’Brien drag model in ANSYS Fluent, MFiX, and OpenFOAM. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 5
<p>Pressure drop for different superficial gas velocities with the Gidaspow drag model in ANSYS Fluent, MFiX, and OpenFOAM. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 6
<p>Time-averaged solid velocities recorded between 3 and 12 s for a superficial gas velocity of 0.38 m/s with Syamlal–O’Brien drag model. These measurements were conducted at a height of y = 0.2 m.</p>
Full article ">Figure 7
<p>Time-averaged solid velocities recorded between 3 and 12 s for a superficial gas velocity of 0.38 m/s with Gidaspow drag model. These measurements were conducted at a height of y = 0.2 m.</p>
Full article ">Figure 8
<p>Time-averaged voidage recorded between 3 and 12 s for a superficial gas velocity of 0.38 m/s with Syamlal–O’Brien drag model. These measurements were conducted at a height of y = 0.2 m. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 9
<p>Time-averaged voidage recorded between 3 and 12 s for a superficial gas velocity of 0.38 m/s with Gidaspow drag model. These measurements were conducted at a height of y = 0.2 m. Experiment data were taken from [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>], with permission from Elsevier.</p>
Full article ">Figure 10
<p>Particle distribution throughout the simulation with the Syamlal–O’Brien drag model with various superficial gas velocities in ANSYS Fluent, MFiX, and OpenFOAM.</p>
Full article ">Figure 11
<p>Particle distribution throughout the simulation with the Gidaspow drag model with various superficial gas velocities in ANSYS Fluent, MFiX, and OpenFOAM.</p>
Full article ">Figure 12
<p>Effect of number of cells to the bed expansion ratio at a superficial gas velocity of 0.38 m/s using the Syamlal–O’Brien drag model in ANSYS Fluent (Time-averaged from 3 to 10.5 s).</p>
Full article ">Figure 13
<p>Simulation Error based on experimental data of Taghipour et al. [<a href="#B19-dynamics-04-00025" class="html-bibr">19</a>] at y = 0.2 and <span class="html-italic">u</span> = 0.38 m/s, with permission from Elsevier.</p>
Full article ">
18 pages, 2910 KiB  
Article
An Optimum Design for a Fast-Response Solenoid Valve: Application to a Limaçon Gas Expander
by Md Shazzad Hossain, Ibrahim Sultan, Truong Phung and Apurv Kumar
Dynamics 2024, 4(2), 457-474; https://doi.org/10.3390/dynamics4020024 - 3 Jun 2024
Cited by 1 | Viewed by 745
Abstract
Organic Rankine Cycle (ORC)–based small-scale power plants are becoming a promising instrument in the recent drive to utilize renewable sources and reduce carbon emissions. But the effectiveness of such systems is limited by the low efficiency of gas expanders, which are the main [...] Read more.
Organic Rankine Cycle (ORC)–based small-scale power plants are becoming a promising instrument in the recent drive to utilize renewable sources and reduce carbon emissions. But the effectiveness of such systems is limited by the low efficiency of gas expanders, which are the main part of an ORC system. Limaçon-based expansion machines with a fast inlet control valve have great prospects as they could potentially offer efficiencies over 50%. However, the lack of a highly reliable and significantly fast control valve is hindering its possible application. In this paper, a push–pull solenoid valve is optimized using a stochastic optimization technique to provide a fast response. The optimization yields about 56–58% improvement in overall valve response. A performance comparison of the initial and optimized valves applied to a limaçon expander thermodynamic model is also presented. Additionally, the sensitivity of the valve towards a changing inlet pressure and expander rotor velocity is analyzed to better understand the effectiveness of the valve and provide clues to overall performance improvement. Full article
Show Figures

Figure 1

Figure 1
<p>Limaçon gas expander with inlet valve.</p>
Full article ">Figure 2
<p>Push–pull solenoid valve.</p>
Full article ">Figure 3
<p>Lever mechanism.</p>
Full article ">Figure 4
<p>Ideal valve operation.</p>
Full article ">Figure 5
<p>Optimization flowchart.</p>
Full article ">Figure 6
<p>Optimization process. (<b>a</b>) Minimization of cost function and (<b>b</b>) response delay.</p>
Full article ">Figure 7
<p>Solenoid (<b>a</b>) current and (<b>b</b>) force.</p>
Full article ">Figure 8
<p>Valve response (<b>a</b>) velocity and (<b>b</b>) area.</p>
Full article ">Figure 9
<p>Fluid leakage.</p>
Full article ">Figure 10
<p>Isentropic efficiency: (<b>a</b>) variation and (<b>b</b>) contours at different rotor velocity and pressure.</p>
Full article ">Figure 11
<p>Isentropic efficiency at different pressure and volume ratios.</p>
Full article ">Figure 12
<p>Filling factor: (<b>a</b>) variation and (<b>b</b>) contours at different rotor velocity and pressure.</p>
Full article ">Figure 13
<p>Air density at different pressures.</p>
Full article ">Figure 14
<p>Mass flow: (<b>a</b>) variation and (<b>b</b>) contours at different rotor velocity and pressure.</p>
Full article ">Figure 15
<p>Energy: (<b>a</b>) variation and (<b>b</b>) contours at different rotor velocity and pressure.</p>
Full article ">Figure 16
<p>Indicated power: (<b>a</b>) variation and (<b>b</b>) contours at different rotor velocity and pressure.</p>
Full article ">
32 pages, 21013 KiB  
Article
Artificial Intelligence Modeling of the Heterogeneous Gas Quenching Process for Steel Batches Based on Numerical Simulations and Experiments
by Nithin Mohan Narayan, Pierre Max Landgraf, Thomas Lampke and Udo Fritsching
Dynamics 2024, 4(2), 425-456; https://doi.org/10.3390/dynamics4020023 - 3 Jun 2024
Viewed by 851
Abstract
High-pressure gas quenching is widely used in the metals industry during the heat treatment processing of steel specimens to improve their material properties. In a gas quenching process, a preheated austenised metal specimen is rapidly cooled with a gas such as nitrogen, helium, [...] Read more.
High-pressure gas quenching is widely used in the metals industry during the heat treatment processing of steel specimens to improve their material properties. In a gas quenching process, a preheated austenised metal specimen is rapidly cooled with a gas such as nitrogen, helium, etc. The resulting microstructure relies on the temporal and spatial thermal history during the quenching. As a result, the corresponding material properties such as hardness are achieved. Challenges reside with the selection of the proper process parameters. This research focuses on the heat treatment of steel sample batches. The gas quenching process is fundamentally investigated in experiments and numerical simulations. Experiments are carried out to determine the heat transfer coefficient and the cooling curves as well as the local flow fields. Quenched samples are analyzed to derive the material hardness. CFD and FEM models numerically determine the conjugate heat transfer, flow behavior, cooling curve, and material hardness. In a novel approach, the experimental and simulation results are adopted to train artificial neural networks (ANNs), which allow us to predict the required process parameters for a targeted material property. The steels 42CrMo4 (1.7225) and 100Cr6 (1.3505) are investigated, nitrogen is the quenching gas, and geometries such as a disc, disc with a hole and ring are considered for batch series production. Full article
Show Figures

Figure 1

Figure 1
<p>Sample geometries for investigation such as disc, disc with hole, and ring; plastic probes for cold experiments.</p>
Full article ">Figure 2
<p>Experimental setup. (<b>a</b>) Suction wind tunnel for measuring HTC. (<b>b</b>) Model gas quenching chamber for investigating flow field with batches (cold experiments). (<b>c</b>) Two-chamber heat treatment plant (Ipsen) for quenching of batches.</p>
Full article ">Figure 3
<p>(<b>a</b>) A two-layered inline batch from disc probes with measurement systems below the second layer for quenching in the Ipsen plant. (<b>b</b>) The probe designation considered in this work within the batch for Ipsen plant experiments. (<b>c</b>) The CAD diagram showing measurement locations within the probe volume for thermocouples; the thermocouple directed from the bottom to top surface. (<b>d</b>) The thermocouple supported by an additional support plate at the bottom surface of the probe to be quenched.</p>
Full article ">Figure 4
<p>(<b>a</b>) The 1D-CTA probe for velocity measurement. (<b>b</b>) The pitot tube for velocity measurement. (<b>c</b>) The film probe (glue-on) from DANTEC [<a href="#B21-dynamics-04-00023" class="html-bibr">21</a>] for determining the HTC. (<b>d</b>) XY positioning stages with the flow measuring probe [<a href="#B22-dynamics-04-00023" class="html-bibr">22</a>] assembled inside the model chamber for analyzing the flow behavior past batches.</p>
Full article ">Figure 5
<p>The pattern of the hardness measurements (indentations) for different probe geometries such as a disc, disc with a hole, and ring at the top and bottom probe surfaces.</p>
Full article ">Figure 6
<p>(<b>a</b>) Flow velocity at wind tunnel center measured with pitot tube and 1D-CTA probe. (<b>b</b>) Measurement repeatability with wind tunnel for HTC at mean flow of 7.8 m/s (16 Hz).</p>
Full article ">Figure 7
<p>HTC measurements for the disc probe at two local positions and for varying mean flow velocity (7.8, 9.8, and 12 m/s).</p>
Full article ">Figure 8
<p>The flow field analysis within the model chamber. (<b>a</b>) The support frame for holding samples and the flow measurement system below the support frame with the xy positioning stage. (<b>b</b>) Sample two-layered inline discs with hole arrangement inside the model chamber. (<b>c</b>) The flow measurement window of 200 × 300 mm<sup>2</sup>.</p>
Full article ">Figure 9
<p>(<b>a</b>) Flow field measurement for an empty model chamber at a distance of 130 mm (inlet position) above the top layer with varying flow rates resp. to motor frequency. (<b>b</b>) Flow field measurement at a distance of 20 mm below the bottom layer for inline two-layered batches with discs, discs with a hole and rings at a motor frequency of 30 Hz resp. to a volumetric flow rate of 3.48 m<sup>3</sup>/s.</p>
Full article ">Figure 10
<p>(<b>a</b>) Measurement reproducibility for 10 bar N<sub>2</sub> quenching measured at the austenitic probe (1.4301/disc) with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min. (<b>b</b>) The quenching trend within an inline disc batch during 10 bar N<sub>2</sub> quenching with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min. (<b>c</b>) Quenching intensity during 10 bar N<sub>2</sub> quenching within a disc probe (42CrMo4) with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min at probe position T-2-2. (<b>d</b>) Quenching intensity during 10 bar N<sub>2</sub> quenching within a ring probe (42CrMo4) with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min at probe position T-2-2.</p>
Full article ">Figure 11
<p>(<b>a</b>) Influence of gas pressure on sample cooling within inline disc batch configuration (42CrMo4); N<sub>2</sub> quenching with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min. (<b>b</b>) Influence of specimen geometry on t<sub>800/500</sub> for inline batch (42CrMo4) during 10 bar N<sub>2</sub> quenching with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min. (<b>c</b>) Influence of specimen material on quenching for inline disc batch (42CrMo4) during N<sub>2</sub> quenching with T<sub>f</sub> = 850 °C and t<sub>f</sub> = 75 min.</p>
Full article ">Figure 12
<p>HRC–hardness of heat-treated specimen at top and bottom surfaces. (<b>a</b>) Disc at position T-1-1 (42CrMo4—10 bar). (<b>b</b>) Disc at position T-1-1 (42CrMo4—6 bar). (<b>c</b>) Ring at position T-1-1 (42CrMo4—10 bar). (<b>d</b>) Ring at position T-2-3 (100Cr6—10 bar).</p>
Full article ">Figure 13
<p>Computational domain and spatial discretization for (<b>a</b>) single-probe CFD simulation; (<b>b</b>) disc batch CFD simulation.</p>
Full article ">Figure 14
<p>(<b>a</b>) Complex interactions in heat treatment simulation using FEM, taking into account microstructural transformation, adapted from [<a href="#B29-dynamics-04-00023" class="html-bibr">29</a>,<a href="#B30-dynamics-04-00023" class="html-bibr">30</a>]. (<b>b</b>,<b>c</b>) Example of selected material data for 42CrMo4 from JMatPro<sup>®</sup>: thermal conductivity and volumetric heat capacity.</p>
Full article ">Figure 15
<p>(<b>a</b>) An example of the structure, input, and output parameters of the ANN, which has been created and trained based on the experimentally determined hardness values. (<b>b</b>) An example of the structure, input and output parameters of the ANN created and trained on the basis of the hardness values determined by CFD and FEM simulations.</p>
Full article ">Figure 16
<p>(<b>a</b>) Validation of numerical model with experiment, HTC for disc probe with V<sub>avg</sub> = 7.8 m/s and T<sub>probe</sub> = 100 °C, and corresponding HTC contours [<a href="#B14-dynamics-04-00023" class="html-bibr">14</a>]. (<b>b</b>) Velocity vector for simulation with V<sub>avg</sub> = 7.8 m/s and T<sub>probe</sub> = 100 °C for disc probe [<a href="#B14-dynamics-04-00023" class="html-bibr">14</a>]. (<b>c</b>) Influence of flow velocity on HTC with T<sub>probe</sub> = 100 °C for disc probe.</p>
Full article ">Figure 17
<p>Simulation results for the ring with V<sub>avg</sub> = 7.8 m/s and T<sub>probe</sub> = 100 °C. (<b>a</b>) The velocity vector; (<b>b</b>) HTC contours.</p>
Full article ">Figure 18
<p>(<b>a</b>) HTC contours and velocity vectors during N<sub>2</sub> quenching with T<sub>probe</sub> = 850 °C (42CrMo4), P = 10 bar and V<sub>avg</sub> = 13.4 m/s. (<b>a</b>) Inline disc batch. (<b>b</b>) Inline ring batch. (<b>c</b>) Staggered disc batch.</p>
Full article ">Figure 19
<p>Influence of (<b>a</b>) gas pressure and (<b>b</b>) flow velocity during inline batch N<sub>2</sub> quenching with T<sub>probe</sub> = 850 °C (42CrMo4).</p>
Full article ">Figure 20
<p>Impact of flow velocity and gas pressure on (<b>a</b>) HTC<sub>max</sub> and (<b>b)</b> HTC<sub>avg</sub> based on correlations developed for disc batch (T-2-2/T-2-3) for 42CrMo4 with T<sub>probe</sub> = 850 °C during N<sub>2</sub> quenching.</p>
Full article ">Figure 21
<p>(<b>a</b>) Comparison of simulated hardness values based on different transformation data with experimentally determined hardness values. (<b>b</b>) Examples of selected FEM results (2D and 3D representation) for T<sub>probe</sub> = 850 °C and P = 10 bar. (<b>c</b>) Comparison of experimentally (Exp) determined cooling curve from sample core with simulated (Sim) cooling curve (FEM) based on the CFD boundary conditions (HTC) for disc, disc with hole, and ring specimen geometries during 10-bar quenching from T<sub>f</sub> = 850 °C.</p>
Full article ">Figure 22
<p>(<b>a</b>) Regression total output ANN with target variable. (<b>b</b>) Differentiation between training, validation and test data. (<b>c</b>) Hardness values of experiment and ANN. (<b>d</b>) Deviating measured values marked by red arrow.</p>
Full article ">Figure 23
<p>(<b>a</b>) Hardness determined experimentally and from ANN for probes’ disc, disc with hole, and ring. (<b>b</b>) Example of calculated hardness distribution in HRC for disc at position T-1-1 (42CrMo4) at gas quenching pressure of 8 bar. (<b>c</b>) Regression total output ANN with target value based on ANN/FEM results.</p>
Full article ">Figure 24
<p>(<b>Top</b>) Differentiation between training, validation, and test data for outputs; (<b>Bottom</b>) output values of the experiment and ANN; outputs 1–5 are the hardness and volume fraction for austenite, pearlite, bainite and martensite.</p>
Full article ">Figure 25
<p>(<b>a</b>) The variation of the gas pressure on the example disc T-1-1 (top surface) from 42CrMo4; the 8-bar deviate forms the resulting image. (<b>b</b>) The variation of the mean values calculated by the ANN as well as the maxima and minima and optimum at the required minimum hardness (for example, 40 HRC requires quenching with min. of 10.25-bar gas pressure); the deviation at 8 bar is also observed here.</p>
Full article ">Figure 26
<p>Comparison of hardness profiles from (<b>a</b>) Ipsen batch heat treatment experiment for disc batch and (<b>b</b>) other industrial chambers for disc batch with 10 bar N<sub>2</sub> quenching; global scale (global maximum hardness for hardness legend) and local scale (local maximum hardness for hardness legend) illustrated for qualitative comparison.</p>
Full article ">
31 pages, 5192 KiB  
Review
Cupolets: History, Theory, and Applications
by Matthew A. Morena and Kevin M. Short
Dynamics 2024, 4(2), 394-424; https://doi.org/10.3390/dynamics4020022 - 13 May 2024
Viewed by 699
Abstract
In chaos control, one usually seeks to stabilize the unstable periodic orbits (UPOs) that densely inhabit the attractors of many chaotic dynamical systems. These orbits collectively play a significant role in determining the dynamics and properties of chaotic systems and are said to [...] Read more.
In chaos control, one usually seeks to stabilize the unstable periodic orbits (UPOs) that densely inhabit the attractors of many chaotic dynamical systems. These orbits collectively play a significant role in determining the dynamics and properties of chaotic systems and are said to form the skeleton of the associated attractors. While UPOs are insightful tools for analysis, they are naturally unstable and, as such, are difficult to find and computationally expensive to stabilize. An alternative to using UPOs is to approximate them using cupolets. Cupolets, a name derived from chaotic, unstable, periodic, orbit-lets, are a relatively new class of waveforms that represent highly accurate approximations to the UPOs of chaotic systems, but which are generated via a particular control scheme that applies tiny perturbations along Poincaré sections. Originally discovered in an application of secure chaotic communications, cupolets have since gone on to play pivotal roles in a number of theoretical and practical applications. These developments include using cupolets as wavelets for image compression, targeting in dynamical systems, a chaotic analog to quantum entanglement, an abstract reducibility classification, a basis for audio and video compression, and, most recently, their detection in a chaotic neuron model. This review will detail the historical development of cupolets, how they are generated, and their successful integration into theoretical and computational science and will also identify some unanswered questions and future directions for this work. Full article
(This article belongs to the Special Issue Recent Advances in Dynamic Phenomena—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Double scroll attractor projected into the <math display="inline"><semantics> <mrow> <msub> <mi>v</mi> <msub> <mi>C</mi> <mn>1</mn> </msub> </msub> <mo>−</mo> <msub> <mi>i</mi> <mi>L</mi> </msub> </mrow> </semantics></math> plane, with control planes intersecting lobes 0 and 1 [<a href="#B11-dynamics-04-00022" class="html-bibr">11</a>]. Reproduced from Kevin M. Short and Matthew A. Morena, “Signatures of quantum mechanics in chaotic systems”, Entropy 21(6), 618 (2019) <a href="https://doi.org/10.3390/e21060618" target="_blank">https://doi.org/10.3390/e21060618</a> (accessed on 9 March 2024), with the permission of MDPI.</p>
Full article ">Figure 2
<p>Comparing the <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> </semantics></math> function for <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <mn>5</mn> </mrow> </semantics></math> (red) future intersections with Poincaré surfaces and <math display="inline"><semantics> <mrow> <mi>P</mi> <mo>=</mo> <mn>12</mn> </mrow> </semantics></math> (blue) future intersections with the Poincaré surfaces.</p>
Full article ">Figure 3
<p>Plot of the simplest cupolet, <b>C</b>00. Phase space plot in left column; time series plots for <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> coordinates in center column; Magnitude FFT spectrum of four periods of the corresponding <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> time series in right column.</p>
Full article ">Figure 4
<p>Plot of cupolet <b>C</b>000101. Phase space plot in left column; time series plots for <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> coordinates in center column; Magnitude FFT spectrum of four periods of the corresponding <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> time series in right column.</p>
Full article ">Figure 5
<p>Plot of cupolet <b>C</b>00101. Phase space plot in left column; time series plots for <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> coordinates in center column; Magnitude FFT spectrum of four periods of the corresponding <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> time series in right column.</p>
Full article ">Figure 6
<p>Plot of cupolet <b>C</b>01011. Phase space plot in left column; time series plots for <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> coordinates in center column; Magnitude FFT spectrum of four periods of the corresponding <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>−</mo> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>−</mo> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>−</mo> </mrow> </semantics></math> time series in right column.</p>
Full article ">Figure 7
<p>Plot of cupolet <b>C</b>01011011. Phase space plot in left column; time series plots for <span class="html-italic">x</span>-, <span class="html-italic">y</span>-, and <span class="html-italic">z</span>-coordinates in center column; Magnitude FFT spectrum of 4 periods of the corresponding <span class="html-italic">x</span>-, <span class="html-italic">y</span>-, and <span class="html-italic">z</span>-time series in right column.</p>
Full article ">Figure 8
<p>Illustrating the diversity in cupolets with (<b>a</b>) their spectral variation among cupolets and (<b>b</b>) their time-domain variations [<a href="#B52-dynamics-04-00022" class="html-bibr">52</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Original <math display="inline"><semantics> <mrow> <mn>256</mn> <mo>×</mo> <mn>256</mn> </mrow> </semantics></math> image and incremental resolution levels. The number of basis elements in each resolution level is (<b>b</b>) 24 (<b>c</b>) 56 and (<b>d</b>) 120 per window [<a href="#B60-dynamics-04-00022" class="html-bibr">60</a>]. Reproduced from Kourosh Zarringhalam and Kevin M. Short, “Generating an adaptive multiresolution image analysis with compact cupolets”, Nonlinear Dynamics, 52, 51-70 (2008) <a href="https://doi.org/10.1007/s11071-007-9257-7" target="_blank">https://doi.org/10.1007/s11071-007-9257-7</a> (accessed on 9 March 2024), with the permission of Springer Publishing.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) Original <math display="inline"><semantics> <mrow> <mn>256</mn> <mo>×</mo> <mn>256</mn> </mrow> </semantics></math> image and incremental resolution levels. The number of basis elements in each resolution level is (<b>b</b>) 24 (<b>c</b>) 56 and (<b>d</b>) 120 per window [<a href="#B60-dynamics-04-00022" class="html-bibr">60</a>]. Reproduced from Kourosh Zarringhalam and Kevin M. Short, “Generating an adaptive multiresolution image analysis with compact cupolets”, Nonlinear Dynamics, 52, 51-70 (2008) <a href="https://doi.org/10.1007/s11071-007-9257-7" target="_blank">https://doi.org/10.1007/s11071-007-9257-7</a> (accessed on 9 March 2024), with the permission of Springer Publishing.</p>
Full article ">Figure 10
<p>(<b>a</b>) Cupolet <b>C</b>00 (blue orbit) destabilizes to a chaotic transient (purple orbit) when the implementation of its control sequence is disrupted, (<b>b</b>) a smooth, transient−free transition from <b>C</b>00 to <b>C</b>01 (red orbit), (<b>c</b>) a lengthy transition from <b>C</b>00 to <b>C</b>01 that involves multiple intermediary loops around the attractor, (<b>d</b>) time series showing the transition seen in (<b>c</b>). Reproduced from Matthew A. Morena, Kevin M. Short, and Erica E. Cooke, “Controlled transitions between cupolets of chaotic systems”, Chaos 24, 013111 (2014) <a href="https://doi.org/10.1063/1.4862668" target="_blank">https://doi.org/10.1063/1.4862668</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 11
<p>A weighted digraph showing the interconnectivity of cupolets as edges and control bins as vertices. Reproduced from Matthew A. Morena, Kevin M. Short, and Erica E. Cooke, “Controlled transitions between cupolets of chaotic systems”, Chaos 24, 013111 (2014) <a href="https://doi.org/10.1063/1.4862668" target="_blank">https://doi.org/10.1063/1.4862668</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 12
<p>Illustrating the amalgamation of composite cupolet <b>C</b>0001111101110111 (period-16) by simpler, fundamental cupolets <b>C</b>000110111 (period-9), <b>C</b>01111 (period-5), and <b>C</b>11 (period-2). Reproduced from Matthew A. Morena and Kevin M. Short, “Fundamental cupolets of chaotic systems”, Chaos, 30, 093114 (2020) <a href="https://doi.org/10.1063/5.0003443" target="_blank">https://doi.org/10.1063/5.0003443</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 13
<p>Step-by-step formation of chaotic entanglement for the double scroll system [<a href="#B11-dynamics-04-00022" class="html-bibr">11</a>]. (<b>a</b>) An external control is applied to System I. (<b>b</b>) This subsequently stabilizes System I onto cupolet <b>C</b>000011111, which then begins producing symbolic dynamics information in the form of its visitation sequence. (<b>c</b>) An interaction function converts the visitation sequence into the emitted sequence <b>E</b>011101111 and transmits this information as a control to System II. This induces System II to stabilize onto cupolet <b>C</b>011101111. (<b>d</b>) Cupolet <b>C</b>011101111’s visitation sequence is then converted to the emitted sequence <b>E</b>000011111 and transmitted as control information back to System I, replacing the external control. Each emitted sequence exactly matches the corresponding cupolet’s control sequence, which locks both systems onto the persistent, periodic orbits of their cupolets. The cupolets’ entanglement will perpetuate as long as the exchange function continues mediating the exchange of control information between the systems. Reproduced from Kevin M. Short and Matthew A. Morena, “Signatures of quantum mechanics in chaotic systems”, Entropy 21(6), 618 (2019) <a href="https://doi.org/10.3390/e21060618" target="_blank">https://doi.org/10.3390/e21060618</a> (accessed on 9 March 2024), with the permission of MDPI.</p>
Full article ">Figure 14
<p>Cupolets (<b>a</b>) <b>C</b>000011 (period 24) and (<b>b</b>) <b>C</b>000111 (period 18) can be induced into chaotic entanglement via an integrate−and−fire interaction function. The cupolets’ visitation and emitted sequences are listed in <a href="#dynamics-04-00022-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 15
<p>Numerical integration of Equation (<a href="#FD5-dynamics-04-00022" class="html-disp-formula">5</a>) with <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>=</mo> <mn>0.006</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>I</mi> <mo>=</mo> <mn>3.25</mn> </mrow> </semantics></math> so that the model is in a chaotic region [<a href="#B93-dynamics-04-00022" class="html-bibr">93</a>]. (<b>A</b>) 3-Dimensional phase space plot of the <span class="html-italic">x</span>, <span class="html-italic">y</span>, and <span class="html-italic">z</span> dynamics. Projection of (<b>A</b>) onto (<b>B</b>) the <span class="html-italic">x</span>-<span class="html-italic">y</span> plane, (<b>C</b>) the <span class="html-italic">x</span>-<span class="html-italic">z</span> plane, and (<b>D</b>) the <span class="html-italic">y</span>-<span class="html-italic">z</span> plane. Reproduced from John E. Parker and Kevin M. Short, “Cupolets in a chaotic neuron model”, Chaos 32, 113104 (2022) <a href="https://doi.org/10.1063/5.0101667" target="_blank">https://doi.org/10.1063/5.0101667</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 16
<p>One dimensional time series of each dynamical variable (<b>A</b>) <span class="html-italic">x</span>, (<b>B</b>) <span class="html-italic">y</span>, and (<b>C</b>) <span class="html-italic">z</span> in <a href="#dynamics-04-00022-f015" class="html-fig">Figure 15</a> and Equation (<a href="#FD5-dynamics-04-00022" class="html-disp-formula">5</a>). Reproduced from John E. Parker and Kevin M. Short, “Cupolets in a chaotic neuron model”, Chaos 32, 113104 (2022) <a href="https://doi.org/10.1063/5.0101667" target="_blank">https://doi.org/10.1063/5.0101667</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 17
<p>Plot of the Hindmarsh-Rose dynamical system in chaotic regime with control planes [<a href="#B93-dynamics-04-00022" class="html-bibr">93</a>]. Reproduced from John E. Parker and Kevin M. Short, “Cupolets in a chaotic neuron model”, Chaos 32, 113104 (2022) <a href="https://doi.org/10.1063/5.0101667" target="_blank">https://doi.org/10.1063/5.0101667</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 18
<p>Selection of Hindmarsh-Rose cupolets: (<b>A</b>) <b>C</b>11, (<b>B</b>) <b>C</b>0110, (<b>C</b>) <b>C</b>1010010, and (<b>D</b>) <b>C</b>01010010 [<a href="#B93-dynamics-04-00022" class="html-bibr">93</a>]. Reproduced from John E. Parker and Kevin M. Short, “Cupolets in a chaotic neuron model”, Chaos 32, 113104 (2022) <a href="https://doi.org/10.1063/5.0101667" target="_blank">https://doi.org/10.1063/5.0101667</a> (accessed on 9 March 2024), with the permission of AIP Publishing.</p>
Full article ">Figure 19
<p>(<b>a</b>–<b>d</b>) Homologous cupolets from the Hindmarsh-Rose system. All four cupolets derive from the control code 11100010, and the resulting cupolet is dependent on the initial condition.</p>
Full article ">
37 pages, 2184 KiB  
Article
Dynamics of Vortex Structures: From Planets to Black Hole Accretion Disks
by Elizabeth P. Tito and Vadim I. Pavlov
Dynamics 2024, 4(2), 357-393; https://doi.org/10.3390/dynamics4020021 - 13 May 2024
Viewed by 749
Abstract
Thermo-vortices (bright spots, blobs, swirls) in cosmic fluids (planetary atmospheres, or even black hole accretion disks) are sometimes observed as clustered into quasi-symmetrical quasi-stationary groups but conceptualized in models as autonomous items. We demonstrate—using the (analytical) Sharp Boundaries Evolution Method and a generic [...] Read more.
Thermo-vortices (bright spots, blobs, swirls) in cosmic fluids (planetary atmospheres, or even black hole accretion disks) are sometimes observed as clustered into quasi-symmetrical quasi-stationary groups but conceptualized in models as autonomous items. We demonstrate—using the (analytical) Sharp Boundaries Evolution Method and a generic model of a thermo-vorticial field in a rotating “thin” fluid layer in a spacetime that may be curved or flat—that these thermo-vortices may be not independent but represent interlinked parts of a single, coherent, multi-petal macro-structure. This alternative conceptualization may influence the designs of numerical models and image-reconstruction methods. Full article
Show Figures

Figure 1

Figure 1
<p><b>First Row, Left Panel</b> (<b>A</b>): Composite image of the black hole Sgr A* derived from radio (1.3 mm) data collected by the Event Horizon Telescope (EHT) Collaboration [<a href="#B1-dynamics-04-00021" class="html-bibr">1</a>]. <b>First Row, Right Panel</b> (<b>B</b>): Multiple cyclones on Jupiter’s North Pole [<a href="#B7-dynamics-04-00021" class="html-bibr">7</a>]. <b>Second Row, Left Panel</b> (<b>C</b>): The Antarctic Ozone Hole in Earth’s stratosphere; the vortex is quasi-circular on 24 September 2001. <b>Second Row, Right Panel</b> (<b>D</b>): The vortex evolved into a two-petal structure by 24 September 2002. Credit: NASA’s Earth Observatory [<a href="#B8-dynamics-04-00021" class="html-bibr">8</a>].</p>
Full article ">Figure 2
<p>Illustration of a vortex blob.</p>
Full article ">Figure 3
<p>Examples of 3-petal and 5-petal contours.</p>
Full article ">Figure 4
<p>Impact of nonlinearity (peak/trough) parameter <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math> on a three-petal thermo-vortex. <b>Left panel</b> (<b>A</b>): Plot for <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>(</mo> <mi>n</mi> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mi>ϵ</mi> <mo>)</mo> </mrow> </semantics></math>. Dots correspond to examples in Panel 3B. Maximum <math display="inline"><semantics> <mi>μ</mi> </semantics></math> is at <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>: <math display="inline"><semantics> <mrow> <msub> <mi>μ</mi> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi>n</mi> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> <mo>=</mo> <msup> <mi>n</mi> <mn>2</mn> </msup> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math>. <b>Right panel</b> (<b>B</b>): Deformation of three-petal contour: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.07</mn> </mrow> </semantics></math> (green; inner); <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.14</mn> </mrow> </semantics></math> (brown); <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.28</mn> </mrow> </semantics></math> (blue); <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.35</mn> </mrow> </semantics></math> (black; outer).</p>
Full article ">Figure 5
<p>Sensitivity of <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>(</mo> <mi>n</mi> <mo>,</mo> <mi>ϵ</mi> <mo>)</mo> </mrow> </semantics></math> to the number of petals <span class="html-italic">n</span> and nonlinearity (peak/trough) parameter <math display="inline"><semantics> <mi>ϵ</mi> </semantics></math>. <b>Left panel</b> (<b>A</b>): Vertical <math display="inline"><semantics> <mi>μ</mi> </semantics></math>-axis is logarithmic. Curves denote the following: <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </mrow> </semantics></math> (brown; upper); <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.28</mn> </mrow> </semantics></math> (black); <math display="inline"><semantics> <mrow> <mi>ϵ</mi> <mo>=</mo> <mo>−</mo> <mn>0.35</mn> </mrow> </semantics></math> (blue; lower). <b>Right panel</b> (<b>B</b>): Vertical <math display="inline"><semantics> <mi>μ</mi> </semantics></math>-axis is linear and cut off at <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math> for better clarity of the image.</p>
Full article ">Figure 6
<p><b>Left panel</b> (<b>A</b>): Vorticity distribution <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>(</mo> <mi mathvariant="bold">x</mi> <mo>)</mo> </mrow> </semantics></math> of the three-petal thermo-vorticial structure. <b>Right Panel</b> (<b>B</b>): Temperature distribution when <math display="inline"><semantics> <mrow> <mi>y</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> and geometrical parameter <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>l</mi> <mo>/</mo> <mi>R</mi> </mrow> </semantics></math> is as follows: <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mi>π</mi> </mrow> </semantics></math> (black; lower curve), <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>2</mn> <mo>/</mo> <mi>π</mi> </mrow> </semantics></math> (brown), and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mi>π</mi> </mrow> </semantics></math> (blue; upper curve).</p>
Full article ">Figure A1
<p>Domains of spacial coordinates <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>r</mi> <mo>,</mo> <mi>z</mi> <mo>)</mo> </mrow> </semantics></math> where the <math display="inline"><semantics> <msub> <mi>m</mi> <mn>00</mn> </msub> </semantics></math>-component of the metric tensor <math display="inline"><semantics> <msub> <mi>m</mi> <mrow> <mi>i</mi> <mi>k</mi> </mrow> </msub> </semantics></math> becomes negative (darkened zones). Calculations are made for the conditional values of parameters: left panel (<b>A</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>3</mn> <mo>,</mo> <mspace width="0.166667em"/> <mi mathvariant="normal">Ω</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>5</mn> </mrow> </semantics></math>; right panel (<b>B</b>) <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>3</mn> <mo>,</mo> <mspace width="0.166667em"/> <mi mathvariant="normal">Ω</mi> <mo>=</mo> <mn>1</mn> <mo>/</mo> <mn>3</mn> </mrow> </semantics></math>.</p>
Full article ">
20 pages, 1353 KiB  
Article
Estimating Spatio-Temporal Building Power Consumption Based on Graph Convolution Network Method
by Georgios Vontzos, Vasileios Laitsos, Avraam Charakopoulos, Dimitrios Bargiotas and Theodoros E. Karakasidis
Dynamics 2024, 4(2), 337-356; https://doi.org/10.3390/dynamics4020020 - 2 May 2024
Viewed by 989
Abstract
Buildings are responsible for around 30% and 42% of the consumed energy at the global and European levels, respectively. Accurate building power consumption estimation is crucial for resource saving. This research investigates the combination of graph convolutional networks (GCNs) and long short-term memory [...] Read more.
Buildings are responsible for around 30% and 42% of the consumed energy at the global and European levels, respectively. Accurate building power consumption estimation is crucial for resource saving. This research investigates the combination of graph convolutional networks (GCNs) and long short-term memory networks (LSTMs) to analyze power building consumption, thereby focusing on predictive modeling. Specifically, by structuring graphs based on Pearson’s correlation and Euclidean distance methods, GCNs are employed to discern intricate spatial dependencies, and LSTM is used for temporal dependencies. The proposed models are applied to data from a multistory, multizone educational building, and they are then compared with baseline machine learning, deep learning, and statistical models. The performance of all models is evaluated using metrics such as the mean absolute error (MAE), mean squared error (MSE), R-squared (R2), and the coefficient of variation of the root mean squared error (CV(RMSE)). Among the proposed computation models, one of the Euclidean-based models consistently achieved the lowest MAE and MSE values, thus indicating superior prediction accuracy. The suggested methods seem promising and highlight the effectiveness of GCNs in improving accuracy and reliability in predicting power consumption. The results could be useful in the planning of building energy policies by engineers, as well as in the evaluation of the energy management of structures. Full article
Show Figures

Figure 1

Figure 1
<p>Euclidean distance computation in a Cartesian space between zones. This figure represents two floor plans, A and B, with the relevant zones of A and B.</p>
Full article ">Figure 2
<p>The LSTM memory unit architecture.</p>
Full article ">Figure 3
<p>The GCN-LSTM model structure.</p>
Full article ">Figure 4
<p>The floor plans on Floors 1–2 (<b>left</b>) and Floors 3–7 (<b>right</b>) [<a href="#B39-dynamics-04-00020" class="html-bibr">39</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) Letter–value plot of the thermal zones. (<b>b</b>) Plot of the mean and median values of each zone.</p>
Full article ">Figure 6
<p>Letter–value plot of the building’s total power.</p>
Full article ">Figure 7
<p>Correlation plot of the power between zones.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>c</b>): The generated adjacency matrix for each computation method. The edges between the nodes and zones are presented with a black color, e.g., for the PCC method, the nodes with values lower than the threshold were not interconnected with the other nodes due to low or negative correlations.</p>
Full article ">Figure 9
<p>The correlation value distribution plots of the PCC (<b>a</b>) and PCCA (<b>b</b>) methods.</p>
Full article ">Figure 10
<p>The generated adjacency matrices for the distance-based computation methods (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 11
<p>Prediction performance comparison of Zones 3, 11, and 33 for metrics MAE, MSE, and <span class="html-italic">R</span><sup>2</sup>.</p>
Full article ">Figure 12
<p>Plot of Zone 3’s actual and predicted power values grouped per prediction horizon.</p>
Full article ">Figure 13
<p>Prediction performance comparison of the total power for the MAE, MSE, and CV(RMSE) metrics.</p>
Full article ">Figure 14
<p>Plot of the building’s total actual and predicted power values grouped per time step horizon.</p>
Full article ">
15 pages, 322 KiB  
Article
Lie Symmetries of the Wave Equation on the Sphere Using Geometry
by Michael Tsamparlis and Aniekan Magnus Ukpong
Dynamics 2024, 4(2), 322-336; https://doi.org/10.3390/dynamics4020019 - 29 Apr 2024
Viewed by 546
Abstract
A semilinear quadratic equation of the form Aij(x)uij=Bi(x,u)ui+F(x,u) defines a metric Aij; therefore, it is [...] Read more.
A semilinear quadratic equation of the form Aij(x)uij=Bi(x,u)ui+F(x,u) defines a metric Aij; therefore, it is possible to relate the Lie point symmetries of the equation with the symmetries of this metric. The Lie symmetry conditions break into two sets: one set containing the Lie derivative of the metric wrt the Lie symmetry generator, and the other set containing the quantities Bi(x,u),F(x,u). From the first set, it follows that the generators of Lie point symmetries are elements of the conformal algebra of the metric Aij, while the second set serves as constraint equations, which select elements from the conformal algebra of Aij. Therefore, it is possible to determine the Lie point symmetries using a geometric approach based on the computation of the conformal Killing vectors of the metric Aij. In the present article, the nonlinear Poisson equation Δguf(u)=0 is studied. The metric defined by this equation is 1 + 2 decomposable along the gradient Killing vector t. It is a conformally flat metric, which admits 10 conformal Killing vectors. We determine the conformal Killing vectors of this metric using a general geometric method, which computes the conformal Killing vectors of a general 1+(n1) decomposable metric in a systematic way. It is found that the nonlinear Poisson equation Δguf(u)=0 admits Lie point symmetries only when f(u)=ku, and in this case, only the Killing vectors are admitted. It is shown that the Noether point symmetries coincide with the Lie point symmetries. This approach/method can be used to study the Lie point symmetries of more complex equations and with more degrees of freedom. Full article
19 pages, 13364 KiB  
Article
Computational Fluid Dynamics Methodology to Estimate the Drag Coefficient of Balls in Rolling Element Bearings
by Yann Marchesse, Christophe Changenet and Fabrice Ville
Dynamics 2024, 4(2), 303-321; https://doi.org/10.3390/dynamics4020018 - 25 Apr 2024
Viewed by 755
Abstract
The emergence of electric vehicles has brought new issues such as the problem of rolling element bearings (REBs) operating at high speeds. Losses due to these components in mechanical transmissions are a key issue and must therefore be taken into account right from [...] Read more.
The emergence of electric vehicles has brought new issues such as the problem of rolling element bearings (REBs) operating at high speeds. Losses due to these components in mechanical transmissions are a key issue and must therefore be taken into account right from the design stage of these systems. Among these losses, the one induced by the motion of rolling elements, known as drag loss, becomes predominant in high-speed REBs. Although an experimental approach is still possible, it is difficult to isolate this loss in order to study it properly. A numerical approach based on CFD is therefore a possible way forward, even if other issues arise. The aim of this article is to study the ability of such an approach to correctly estimate the drag coefficient associated with the motion of rolling elements. The influence of the numerical domain extension, the mesh refinement, the simplification of the ring shape, and the presence of the cage on the values of the drag coefficient is presented. While it seems possible to compromise on the calculation domain and mesh size, it appears that the other parameters must be taken into account as much as possible to obtain realistic results. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the hypothesis and the simplifications made for the CFD approach.</p>
Full article ">Figure 2
<p>Detailed view of the mesh in the ball, the cage and the inner race and in a plane perpendicular to the flow and definition of the cell size on the ball, <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>∆</mo> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Boundary conditions when one ball is considered in the numerical domain.</p>
Full article ">Figure 4
<p>Time evolution of the drag coefficient (<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </mrow> <mo>=</mo> <mn>1.66</mn> <mo>;</mo> <mo> </mo> <mi>N</mi> <mo>×</mo> <msub> <mrow> <mi>d</mi> </mrow> <mrow> <mi>m</mi> </mrow> </msub> <mo>=</mo> <mn>8.01</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>5</mn> </mrow> </msup> </mrow> </semantics></math>; three balls in the computational domain; three elements in the gap between the balls and the rings; <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>Δ</mo> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <mn>21</mn> </mrow> </mrow> </mrow> </semantics></math>).</p>
Full article ">Figure 5
<p>Streamlines in the plane located in the middle of the REB (<b>a</b>) and pressure coefficient distribution on balls and streamlines in the cage region (<b>b</b>) (The white arrow indicates the direction of rotation).</p>
Full article ">Figure 6
<p>Meshes based on size cells equal at maximum to <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>∆</mo> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <mn>41</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>∆</mo> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <mn>21</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>right</b>); definition of the cell size, <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>∆</mo> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Numerical domain comprising one ball (<b>left</b>), three balls (<b>centre</b>) and five balls (<b>right</b>).</p>
Full article ">Figure 8
<p>Pressure distribution on the ball located at the centre of all the balls when the numerical domain comprises one ball (<b>left</b>), three balls (<b>centre</b>) and five balls (<b>right</b>).</p>
Full article ">Figure 9
<p>Surface streamlines in the plane in the middle of the REB obtained from computational domain comprising one ball (<b>left</b>), three balls (<b>centre</b>) and five balls (<b>right</b>).</p>
Full article ">Figure 10
<p>Numerical domain representing the following: 10 balls, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.331</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>a</b>); 9 balls, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.479</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>b</b>); 7 balls, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.902</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>c</b>); 6 balls, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>2.219</mn> </mrow> </mrow> </mrow> </semantics></math> (<b>d</b>). The configuration with 8 balls, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.66</mn> </mrow> </mrow> </mrow> </semantics></math>, is presented in the centre of <a href="#dynamics-04-00018-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 11
<p>Influence of the relative distance between two consecutive balls on the pressure coefficient distribution ((<b>left</b>), <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </mrow> <mo>=</mo> <mn>1.331</mn> </mrow> </semantics></math>; (<b>centre</b>), <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </mrow> <mo>=</mo> <mn>1.479</mn> </mrow> </semantics></math>; (<b>right</b>), <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </mrow> <mo>=</mo> <mn>2.219</mn> </mrow> </semantics></math>). The white arrow indicates the direction of rotation.</p>
Full article ">Figure 12
<p>Influence of the relative distance between two consecutive balls on the pressure coefficient distribution around the ball ((<b>a</b>), circle perpendicular to the bearing axis; (<b>b</b>), circle aligned with the rotation axis located at <math display="inline"><semantics> <mrow> <mn>0.25</mn> <mo>×</mo> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> from the centre of the ball; the black arrow indicates the direction of rotation).</p>
Full article ">Figure 13
<p>Influence of the relative distance between two consecutive balls on the drag coefficient.</p>
Full article ">Figure 14
<p>Influence of the type of the cage on the pressure coefficient distribution (<math display="inline"><semantics> <mrow> <mi>L</mi> <mo>=</mo> <mn>1.66</mn> <mo>×</mo> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 15
<p>Influence of the cage type on the pressure coefficient distribution around the ball ((<b>a</b>), circle perpendicular to the bearing axis; (<b>b</b>), circle aligned with the rotation axis located at <math display="inline"><semantics> <mrow> <mn>0.25</mn> <mo>×</mo> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> from the centre of the ball).</p>
Full article ">Figure 16
<p>Influence of the thickness, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>t</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>, of the cage on the pressure coefficient on the ball.</p>
Full article ">Figure 17
<p>Influence of the thickness of the cage on the pressure coefficient distribution around the ball ((<b>a</b>), circle perpendicular to the bearing axis; (<b>b</b>), circle aligned with the rotation axis located at <math display="inline"><semantics> <mrow> <mn>0.25</mn> <mo>×</mo> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> </mrow> </semantics></math> from the centre of the ball).</p>
Full article ">Figure 18
<p>Influence of the thickness of the cage on the drag coefficient (<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.66</mn> </mrow> </mrow> </mrow> </semantics></math>).</p>
Full article ">Figure 19
<p>Influence of the rotational speed on the drag coefficient (<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>L</mi> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>D</mi> </mrow> <mrow> <mi>b</mi> </mrow> </msub> <mo>=</mo> <mn>1.66</mn> </mrow> </mrow> </mrow> </semantics></math>).</p>
Full article ">
16 pages, 8922 KiB  
Article
SPH Simulation of Molten Metal Flow Modeling Lava Flow Phenomena with Solidification
by Shingo Tomita, Joe Yoshikawa, Makoto Sugimoto, Hisaya Komen and Masaya Shigeta
Dynamics 2024, 4(2), 287-302; https://doi.org/10.3390/dynamics4020017 - 19 Apr 2024
Viewed by 937
Abstract
Characteristic dynamics in lava flows, such as the formation processes of lava levees, toe-like tips, and overlapped structures, were reproduced successfully through numerical simulation using the smoothed particle hydrodynamics (SPH) method. Since these specific phenomena have a great influence on the flow direction [...] Read more.
Characteristic dynamics in lava flows, such as the formation processes of lava levees, toe-like tips, and overlapped structures, were reproduced successfully through numerical simulation using the smoothed particle hydrodynamics (SPH) method. Since these specific phenomena have a great influence on the flow direction of lava flows, it is indispensable to elucidate them for accurate predictions of areas where lava strikes. At the first step of this study, lava was expressed using a molten metal with known physical properties. The computational results showed that levees and toe-like tips formed at the fringe of the molten metal flowing down on a slope, which appeared for actual lava flows as well. The dynamics of an overlapped structure formation were also simulated successfully; therein, molten metal flowed down, solidified, and changed the surface shape of the slope, and the second molten metal flowed over the changed surface shape. It was concluded that the computational model developed in this study using the SPH method is applicable for simulating and clarifying lava flow phenomena. Full article
(This article belongs to the Special Issue Recent Advances in Dynamic Phenomena—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic of cross-section of lava flow.</p>
Full article ">Figure 2
<p>Computational domain.</p>
Full article ">Figure 3
<p>The temperature distribution of gas at <math display="inline"><semantics> <mrow> <mi>z</mi> <mo>=</mo> <mn>5.1</mn> <mtext> </mtext> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">m</mi> </mrow> </semantics></math> [<a href="#B32-dynamics-04-00017" class="html-bibr">32</a>].</p>
Full article ">Figure 4
<p>The heat input distribution for generating molten metal, which is expressed as ion recombination under arc plasma conditions [<a href="#B31-dynamics-04-00017" class="html-bibr">31</a>].</p>
Full article ">Figure 5
<p>Flow and solidification processes of molten metal with time evolution: (<b>a</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.80</mn> </mrow> </semantics></math> s; (<b>b</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>2.45</mn> </mrow> </semantics></math> s; (<b>c</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.70</mn> </mrow> </semantics></math> s; (<b>d</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.97</mn> </mrow> </semantics></math> s; (<b>e</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>7.08</mn> </mrow> </semantics></math> s; (<b>f</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.49</mn> </mrow> </semantics></math> s; (<b>g</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.55</mn> </mrow> </semantics></math> s; (<b>h</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.73</mn> </mrow> </semantics></math> s; (<b>i</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.01</mn> </mrow> </semantics></math> s; (<b>j</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.99</mn> </mrow> </semantics></math> s.</p>
Full article ">Figure 5 Cont.
<p>Flow and solidification processes of molten metal with time evolution: (<b>a</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.80</mn> </mrow> </semantics></math> s; (<b>b</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>2.45</mn> </mrow> </semantics></math> s; (<b>c</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.70</mn> </mrow> </semantics></math> s; (<b>d</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.97</mn> </mrow> </semantics></math> s; (<b>e</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>7.08</mn> </mrow> </semantics></math> s; (<b>f</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.49</mn> </mrow> </semantics></math> s; (<b>g</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.55</mn> </mrow> </semantics></math> s; (<b>h</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>9.73</mn> </mrow> </semantics></math> s; (<b>i</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.01</mn> </mrow> </semantics></math> s; (<b>j</b>) at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.99</mn> </mrow> </semantics></math> s.</p>
Full article ">Figure 6
<p>Molten metal flow at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.97</mn> </mrow> </semantics></math> s: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> </mrow> </semantics></math>-directional velocity distribution; (<b>b</b>) temperature distribution.</p>
Full article ">Figure 7
<p>Molten metal flow at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.01</mn> </mrow> </semantics></math> s: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> </mrow> </semantics></math>-directional velocity distribution; (<b>b</b>) temperature distribution.</p>
Full article ">Figure 8
<p>Cross-sections at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>18.0</mn> </mrow> </semantics></math> mm and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>5.97</mn> </mrow> </semantics></math> s: (<b>a</b>) phase of particles; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> </mrow> </semantics></math>-directional velocity distribution; (<b>c</b>) temperature distribution.</p>
Full article ">Figure 9
<p>Cross-sections at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>20.0</mn> </mrow> </semantics></math> mm and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.01</mn> </mrow> </semantics></math> s: (<b>a</b>) phase of particles; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> </mrow> </semantics></math>-directional velocity distribution; (<b>c</b>) temperature distribution.</p>
Full article ">Figure 9 Cont.
<p>Cross-sections at <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>=</mo> <mn>20.0</mn> </mrow> </semantics></math> mm and <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>10.01</mn> </mrow> </semantics></math> s: (<b>a</b>) phase of particles; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>x</mi> </mrow> </semantics></math>-directional velocity distribution; (<b>c</b>) temperature distribution.</p>
Full article ">
15 pages, 659 KiB  
Article
Auto-Correlation Functions of Chaotic Binary Sequences Obtained by Alternating Two Binary Functions
by Akio Tsuneda
Dynamics 2024, 4(2), 272-286; https://doi.org/10.3390/dynamics4020016 - 16 Apr 2024
Viewed by 680
Abstract
This paper discusses the auto-correlation functions of chaotic binary sequences obtained by a one-dimensional chaotic map and two binary functions. The two binary functions are alternately used to obtain a binary sequence from a chaotic real-valued sequence. We consider two similar methods and [...] Read more.
This paper discusses the auto-correlation functions of chaotic binary sequences obtained by a one-dimensional chaotic map and two binary functions. The two binary functions are alternately used to obtain a binary sequence from a chaotic real-valued sequence. We consider two similar methods and give the theoretical auto-correlation functions of the new binary sequences, which are expressed by the auto-/cross-correlation functions of the two chaotic binary sequences generated by a single binary function. Furthermore, some numerical experiments are performed to confirm the validity of the theoretical auto-correlation functions. Full article
Show Figures

Figure 1

Figure 1
<p>Bernoulli map and binary functions.</p>
Full article ">Figure 2
<p>Normalized auto-/cross-correlation functions of <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mi>B</mi> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>2</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mi>B</mi> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math>.</p>
Full article ">Figure 3
<p>Piecewise linear map with three sections and binary functions.</p>
Full article ">Figure 4
<p>Auto-/cross-correlation functions of <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>3</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mrow> <mi>P</mi> <mi>L</mi> </mrow> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>4</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mrow> <mi>P</mi> <mi>L</mi> </mrow> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math>.</p>
Full article ">Figure 5
<p>Tent map and binary functions.</p>
Full article ">Figure 6
<p>Auto-/cross-correlation functions of <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>5</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mrow> <mi>T</mi> </mrow> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> and <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msub> <mi>b</mi> <mn>6</mn> </msub> <mrow> <mo>(</mo> <msubsup> <mi>τ</mi> <mrow> <mi>T</mi> </mrow> <mi>n</mi> </msubsup> <mrow> <mo>(</mo> <mi>x</mi> <mo>)</mo> </mrow> <mo>,</mo> <mi>d</mi> <mo>)</mo> </mrow> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math><math display="inline"><semantics> <mrow> <mo>(</mo> <mi>d</mi> <mo>=</mo> <mn>0.15</mn> <mo>,</mo> <mspace width="0.166667em"/> <mn>0.35</mn> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Normalized auto-correlation functions of new binary sequences <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msubsup> <mi>D</mi> <mi>n</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> in Method 1.</p>
Full article ">Figure 7 Cont.
<p>Normalized auto-correlation functions of new binary sequences <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msubsup> <mi>D</mi> <mi>n</mi> <mrow> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </msubsup> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> in Method 1.</p>
Full article ">Figure 8
<p>Normalized auto-correlation functions of new binary sequences <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msubsup> <mi>D</mi> <mi>n</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> in Method 2.</p>
Full article ">Figure 8 Cont.
<p>Normalized auto-correlation functions of new binary sequences <math display="inline"><semantics> <msubsup> <mrow> <mo>{</mo> <msubsup> <mi>D</mi> <mi>n</mi> <mrow> <mo>(</mo> <mn>2</mn> <mo>)</mo> </mrow> </msubsup> <mo>}</mo> </mrow> <mrow> <mi>n</mi> <mo>=</mo> <mn>0</mn> </mrow> <mo>∞</mo> </msubsup> </semantics></math> in Method 2.</p>
Full article ">
18 pages, 5288 KiB  
Article
System Identification Using Self-Adaptive Filtering Applied to Second-Order Gradient Materials
by Thomas Kletschkowski
Dynamics 2024, 4(2), 254-271; https://doi.org/10.3390/dynamics4020015 - 7 Apr 2024
Viewed by 647
Abstract
For many engineering applications, it is sufficient to use the concept of simple materials. However, higher gradients of the kinematic variables are taken into account to model materials with internal length scales as well as to describe localization effects using gradient theories in [...] Read more.
For many engineering applications, it is sufficient to use the concept of simple materials. However, higher gradients of the kinematic variables are taken into account to model materials with internal length scales as well as to describe localization effects using gradient theories in finite plasticity or fluid mechanics. In many approaches, length scale parameters have been introduced that are related to a specific micro structure. An alternative approach is possible, if a thermodynamically consistent framework is used for material modeling, as shown in the present contribution. However, even if sophisticated and thermodynamically consistent material models can be established, there are still not yet standard experiments to determine higher order material constants. In order to contribute to this ongoing discussion, system identification based on the method of self-adaptive filtering is proposed in this paper. To evaluate the effectiveness of this approach, it has been applied to second-order gradient materials considering longitudinal vibrations. Based on thermodynamically consistent models that have been solved numerically, it has been possible to prove that system identification based on self-adaptive filtering can be used effectively for both narrow-band and broadband signals in the field of second-order gradient materials. It has also been found that the differences identified for simple materials and gradient materials allow for condition monitoring and detection of gradient effects in the material behavior. Full article
(This article belongs to the Special Issue Recent Advances in Dynamic Phenomena—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Coordinate system and schematic configuration of grid points.</p>
Full article ">Figure 2
<p>Behavior of simple material. (<b>Left</b>): impulse response. (<b>Right</b>): frequency response.</p>
Full article ">Figure 3
<p>Behavior of gradient material. (<b>Left</b>): impulse response. (<b>Right</b>): frequency response.</p>
Full article ">Figure 4
<p>Time-harmonic excitation applied to simple material. (<b>Left</b>): system response and steady-state solution. (<b>Right</b>): system response and fully identified model of system response.</p>
Full article ">Figure 5
<p>Adaption process of filter for simple material. (<b>Left</b>): learning curve. (<b>Right</b>): development of filter coefficients.</p>
Full article ">Figure 6
<p>Time-harmonic excitation applied to second-order gradient material. (<b>Left</b>): system response and steady-state solution. (<b>Right</b>): system response and fully identified model of system response.</p>
Full article ">Figure 7
<p>Adaption process of filter for second-order material. (<b>Left</b>): Learning curve. (<b>Right</b>): development of filter coefficients.</p>
Full article ">Figure 8
<p>Random excitation applied to simple material. (<b>Left</b>): time domain response. (<b>Right</b>): system response and fully identified model of system response.</p>
Full article ">Figure 9
<p>Adaption process of filter for simple material. (<b>Left</b>): learning curve. (<b>Right</b>): development of two filter coefficients.</p>
Full article ">Figure 10
<p>Random excitation applied to second-order gradient material. (<b>Left</b>): time domain response. (<b>Right</b>): system response and fully identified model of system response.</p>
Full article ">Figure 11
<p>Adaption process of filter for second-order material. (<b>Left</b>): learning curve. (<b>Right</b>): development of two filter coefficients.</p>
Full article ">Figure 12
<p>Normalized filter weights. (<b>Left</b>): simple material. (<b>Right</b>): second-order gradient material.</p>
Full article ">
21 pages, 74549 KiB  
Article
A Versatile Deposition Model for Natural and Processed Surfaces
by Cihan Ates, Rainer Koch and Hans-Jörg Bauer
Dynamics 2024, 4(2), 233-253; https://doi.org/10.3390/dynamics4020014 - 30 Mar 2024
Viewed by 929
Abstract
This paper introduces a robust deposition model designed for exploring the growth dynamics of deposits on surfaces under practical conditions. The study addresses the challenge of characterizing the intricate morphology of deposits, exhibiting significant visual variations. A generative approach is deployed to create [...] Read more.
This paper introduces a robust deposition model designed for exploring the growth dynamics of deposits on surfaces under practical conditions. The study addresses the challenge of characterizing the intricate morphology of deposits, exhibiting significant visual variations. A generative approach is deployed to create diverse natural and engineered surface textures, governed by probabilistic principles. The model’s formulation addresses key questions related to deposition initiation, nucleation point behaviour, spatial scaling, deposit growth rates, spread dynamics, and surface mobility. A versatile algorithm, relying on six parameters and employing nested loops and Gaussian sampling, is developed. The algorithm’s efficacy is examined through extensive simulations, involving variations in nucleation scaling densities, aggregate scaling scenarios, spread factors, and diffusion rates. Surface statistics are computed for simulated deposits and analyzed using Fast Fourier Transform (FFT). The resulting database enables quantitative comparisons of surfaces generated with different parameters, where the database-derived parallel coordinates offer guidance for selecting optimal model parameters to achieve desired surface morphologies. The proposed approach is validated against urea-derived deposits, exhibiting statistical consistency and agreement with experimental observations. Overall, the model’s adaptable framework holds promise for understanding and predicting deposit growth on surfaces in diverse practical scenarios. Full article
Show Figures

Figure 1

Figure 1
<p>Proposed top-down approach to create surfaces with different target properties.</p>
Full article ">Figure 2
<p>Algorithm of the proposed deposit formation model.</p>
Full article ">Figure 3
<p>Details of the model parameters to change the deposit morphology at macroscopic (<b>a</b>) and microscopic (<b>b</b>) scale. Formation of deposit clusters for a set of parameters including surface diffusion is also depicted (<b>c</b>).</p>
Full article ">Figure 4
<p>Visualization of the target library with the model feature space via parallel coordinates. The RMS axis is ranged up to 150 μm for better visibility. The data, interactive html file and plotting script can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S2</a> to generate alternative views.</p>
Full article ">Figure 5
<p>Microscopic view of urea deposits at different magnifications.</p>
Full article ">Figure 6
<p>Impact of cluster type on the generated deposit height profiles with 1 μm resolution over a substrate of 1000 × 1000 units. The remaining parameters are given at the top of each figure groups. The legends show the height profile. The remaining surface profiles for other parameter combinations can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S3</a>.</p>
Full article ">Figure 7
<p>Impact of cluster type on the surface roughness with 1 μm resolution over a substrate of 1000 × 1000 units. <span class="html-italic">X</span> axis gives the calculated RMS for each 2D cut-plane over the 3D deposit topology. Each curve represents a different smoothing factor (SF). The remaining parameters are given at the top of each figure groups. The remaining profiles for other parameter combinations can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S3</a>.</p>
Full article ">Figure 8
<p>Impact of cluster type on the PSD curves with 1 μm resolution over a substrate of 1000 × 1000 units. Each curve represents a different smoothing factor (SF). The remaining parameters are given at the top of each figure groups. The remaining profiles for other parameter combinations can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S3</a>.</p>
Full article ">Figure 9
<p>Impact of surface diffusion on the generated deposit height profiles with 1 μm resolution over a substrate of 1000 × 1000 units. The remaining parameters are given at the top of each figure groups. The legends show the height profile. The remaining surface profiles for other parameter combinations can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S3</a>.</p>
Full article ">Figure 10
<p>Porous structures generated with the proposed algorithm with varying (<b>a</b>) global porosity, (<b>b</b>) pore cluster shape/aspect ratio and (<b>c</b>) pore size distributions.</p>
Full article ">Figure 11
<p>Porous structures generated with the proposed algorithm with macro and micro pore types. Macro-pores are sampled as nucleation events, while micro-pores are generated by sampling the particle type during a RD event for a given nucleus. An example of the pore sampling procedure can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S1</a>.</p>
Full article ">Figure 12
<p>Identification of model parameters with the measured RMS and PSD slope values. The interactive html file and plotting script can be found in <a href="#app1-dynamics-04-00014" class="html-app">Supplementary Material S2</a> to generate alternative views.</p>
Full article ">Figure 13
<p>Surface properties of the generated deposits with <math display="inline"><semantics> <mrow> <mi>S</mi> <mi>F</mi> <mo>=</mo> <mn>0.618</mn> <mo>,</mo> <msub> <mi>R</mi> <mi>i</mi> </msub> <mo>=</mo> <mn>3</mn> <mo>,</mo> <mi>G</mi> <mi>r</mi> <mi>o</mi> <mi>u</mi> <mi>p</mi> <mo>=</mo> <msub> <mi>C</mi> <mn>1</mn> </msub> <mo>,</mo> <mi>α</mi> <mo>=</mo> <mn>0.75</mn> </mrow> </semantics></math>. (<b>a</b>) 2D RMS profiles along the y axis, (<b>b</b>) mean PSD curve for each deposit surface, (<b>c</b>) variations in 3D RMS of the deposits, (<b>d</b>) variations in the slopes of the PSD curves.</p>
Full article ">
11 pages, 517 KiB  
Article
Dynamics and Stability of Double-Walled Carbon Nanotube Cantilevers Conveying Fluid in an Elastic Medium
by Vassil M. Vassilev and Galin S. Valchev
Dynamics 2024, 4(2), 222-232; https://doi.org/10.3390/dynamics4020013 - 27 Mar 2024
Viewed by 2056
Abstract
The paper concerns the dynamics and stability of double-walled carbon nanotubes conveying fluid. The equations of motion adopted in the current study to describe the dynamics of such nano-pipes stem from the classical Bernoulli–Euler beam theory. Several additional terms are included in the [...] Read more.
The paper concerns the dynamics and stability of double-walled carbon nanotubes conveying fluid. The equations of motion adopted in the current study to describe the dynamics of such nano-pipes stem from the classical Bernoulli–Euler beam theory. Several additional terms are included in the basic equations in order to take into account the influence of the conveyed fluid, the impact of the surrounding medium and the effect of the van der Waals interaction between the inner and outer single-walled carbon nanotubes constituting a double-walled one. In the present work, the flow-induced vibrations of the considered nano-pipes are studied for different values of the length of the pipe, its inner radius, the characteristics of the ambient medium and the velocity of the fluid flow, which is assumed to be constant. The critical fluid flow velocities are obtained at which such a cantilevered double-walled carbon nanotube embedded in an elastic medium loses stability. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of an initially straight DWCNT of length <span class="html-italic">L</span> conveying fluid flowing with constant velocity <span class="html-italic">U</span> and geometric characteristics of its cross-section. Here, <span class="html-italic">h</span> is the wall thickness of the tubes, <math display="inline"><semantics> <mrow> <mi>δ</mi> <mo>=</mo> <mi>h</mi> </mrow> </semantics></math> is the initial distance between them, and <math display="inline"><semantics> <msub> <mi>R</mi> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>R</mi> <mn>2</mn> </msub> </semantics></math> are the inside radii of the inner and outer tubes, respectively.</p>
Full article ">Figure 2
<p>The critical flow velocity <math display="inline"><semantics> <msub> <mi>v</mi> <mrow> <mi>c</mi> <mi>r</mi> <mi>i</mi> <mi>t</mi> <mi>i</mi> <mi>c</mi> <mi>a</mi> <mi>l</mi> </mrow> </msub> </semantics></math> of an initially straight cantilevered DWCNT conveying fluid for three values, 0, 50 and 100, of the spring constant <math display="inline"><semantics> <mi>κ</mi> </semantics></math>, and two value of Young’s modulus <span class="html-italic">E</span> of the pipe: <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> TPa (<b>a</b>) and <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> TPa (<b>c</b>). The evolution of the intertube interaction coefficient <math display="inline"><semantics> <msub> <mi>γ</mi> <mn>1</mn> </msub> </semantics></math> with the ratio <math display="inline"><semantics> <mrow> <mi>L</mi> <mo>/</mo> <msub> <mi>R</mi> <mn>1</mn> </msub> </mrow> </semantics></math> of the length of the nano-pipe <span class="html-italic">L</span> to the inside radius of the inner tube <math display="inline"><semantics> <msub> <mi>R</mi> <mn>1</mn> </msub> </semantics></math> for <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> TPa (<b>b</b>) and <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> TPa (<b>d</b>).</p>
Full article ">Figure 3
<p>Profiles of the critical flow velocities of two fluid-conveying DWCNT cantilevers with Young’s moduli of <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> TPa and <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> TPa represented by solid and dashed lines, respectively, in cases when the impact of the ambient elastic medium is characterised by spring constants of <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (<b>left</b>) and <math display="inline"><semantics> <mrow> <mi>κ</mi> <mo>=</mo> <mn>100</mn> </mrow> </semantics></math> (<b>right</b>).</p>
Full article ">
14 pages, 6212 KiB  
Article
Exploiting Domain Partition in Response Function-Based Dynamic Surrogate Modeling: A Continuous Crystallizer Study
by Alessandro Di Pretoro, Ludovic Montastruc and Stéphane Negny
Dynamics 2024, 4(2), 208-221; https://doi.org/10.3390/dynamics4020012 - 26 Mar 2024
Viewed by 585
Abstract
Given the exponential rise in the amount of data requiring processing in all engineering fields, phenomenological models have become computationally cumbersome. For this reason, more efficient data-driven models have been recently used with the purpose of substantially reducing simulation computational times. However, especially [...] Read more.
Given the exponential rise in the amount of data requiring processing in all engineering fields, phenomenological models have become computationally cumbersome. For this reason, more efficient data-driven models have been recently used with the purpose of substantially reducing simulation computational times. However, especially in process engineering, the majority of the proposed surrogate models address steady-state problems, while poor studies refer to dynamic simulation modeling. For this reason, using a response function-based approach, a crystallization unit case study was set up in order to derive a dynamic data-driven model for crystal growth whose characteristic differential parameters are derived via Response Surface Methodology. In particular, multiple independent variables were considered, and a well-established sampling technique was exploited for sample generation. Then, different sample sizes were tested and compared in terms of accuracy indicators. Finally, the domain partition strategy was exploited in order to show its relevant impact on the final model accuracy. In conclusion, the outcome of this study proved that the proposed procedure is a suitable methodology for dynamic system metamodeling, as it shows good compliance and relevant improvement in terms of computational time. In terms of future research perspectives, testing the proposed procedure on different systems and in other research fields would allow for greater improvement and would, eventually, extend its validity. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of evaporative crystallizer.</p>
Full article ">Figure 2
<p>Latin Hypercube Sampling example: (<b>a</b>) 2D space; (<b>b</b>) 3D space.</p>
Full article ">Figure 3
<p>Dynamic profiles for the DoE: (<b>a</b>) independent input variables; (<b>b</b>) crystal growth.</p>
Full article ">Figure 4
<p>Dynamic profiles for the model validation: (<b>a</b>) independent input variables; (<b>b</b>) crystal growth.</p>
Full article ">Figure 5
<p>Dynamic surrogate model results with a single set of characteristic parameters.</p>
Full article ">Figure 6
<p>Comparison between phenomenological and domain partition-wise surrogate model with respect to the DoE dataset: (<b>a</b>) crystal growth ratio; (<b>b</b>) crystal linear growth.</p>
Full article ">Figure 7
<p>Computational performance indicators: (<b>a</b>) (1-NMSE) vs. sample size; (<b>b</b>) computational time vs. sample size.</p>
Full article ">Figure 8
<p>Comparison between phenomenological and domain partition-wise surrogate model with respect to the validation dataset: (<b>a</b>) crystal growth ratio; (<b>b</b>) crystal linear growth.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop