[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

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
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (765)

Search Parameters:
Keywords = low hysteresis

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 4207 KiB  
Article
Predicting Suspended Sediment Transport in Urbanised Streams: A Case Study of Dry Creek, South Australia
by Tesfa Gebrie Andualem, Guna A. Hewa, Baden R. Myers, John Boland and Stefan Peters
Hydrology 2024, 11(11), 196; https://doi.org/10.3390/hydrology11110196 (registering DOI) - 16 Nov 2024
Viewed by 56
Abstract
Sediment transport in urban streams is a critical environmental issue, with significant implications for water quality, ecosystem health, and infrastructure management. Accurately estimating suspended sediment concentration (SSC) is essential for effective long-term environmental management. This study investigates the relationships between streamflow, turbidity, and [...] Read more.
Sediment transport in urban streams is a critical environmental issue, with significant implications for water quality, ecosystem health, and infrastructure management. Accurately estimating suspended sediment concentration (SSC) is essential for effective long-term environmental management. This study investigates the relationships between streamflow, turbidity, and SSC in Dry Creek, South Australia, to understand sediment transport dynamics in urbanised catchments. We collected grab samples from the field and analysed them in the laboratory. We employed statistical modelling to develop a sediment rating curve (SRC) that provides insights into the sediment transport dynamics in the urban stream. The grab sample measurements showed variations in SSC between 3.2 and 431.8 mg/L, with a median value of 77.3 mg/L. The analysis revealed a strong linear relationship between streamflow and SSC, while turbidity exhibited a two-regime linear relationship, in which the low-turbidity regime demonstrated a stronger linear relationship compared to the high-turbidity regime. This is attributed to the urbanised nature of the catchment, which contributes to a first-flush effect in turbidity. This contributes to sediment hysteresis, resulting in non-proportional turbidity and SSC responses to streamflow changes. The findings demonstrate the effectiveness of a streamflow-based SRC for accurately predicting sediment discharge, explaining 97% of the variability in sediment discharge. The sediment discharge predicted using the SRC indicated a sediment load of 341.8 tonnes per year along the creek. The developed sediment rating curve provides a valuable tool for long-term sediment management in Dry Creek, enabling the assessment of downstream environmental risks. By addressing data limitations, this study contributes to a deeper understanding of sediment transport dynamics in urbanized environments, offering insights for informed decision-making and effective sediment management strategies. Full article
(This article belongs to the Special Issue Sediment Transport and Morphological Processes at the Watershed Scale)
Show Figures

Figure 1

Figure 1
<p>Key map and location of sampling site (Conway Crescent Valley View gauging station), distribution of land cover, elevation, and slope within the Dry Creek catchment.</p>
Full article ">Figure 2
<p>The relationship between measured turbidity (turbidity measured at the laboratory) and telemetry turbidity measured online at the Conway Crescent Valley View gauging station (<b>a</b>) with outliers, and (<b>b</b>) without outliers.</p>
Full article ">Figure 3
<p>Continuous streamflow and turbidity data collected via telemetry at the Conway Crescent Valley View gauging station during selected grab sampling events revealing distinct turbidity patterns, (<b>a</b>) first flush effect, (<b>b</b>) complex pattern, and (<b>c</b>) counter-clockwise flow-turbidity relationships.</p>
Full article ">Figure 4
<p>Scatterplots illustrating the relationships between, (<b>a</b>) suspended sediment concentration (SSC) and turbidity (T), (<b>b</b>) SSC and streamflow (Q) in Dry Creek.</p>
Full article ">Figure 5
<p>Comparison of measured and predicted suspended sediment concentration (SSC) using Q-based and T-based models in Dry Creek.</p>
Full article ">Figure 6
<p>Sediment transport dynamics in Dry Creek; (<b>a</b>) sediment rating curve, (<b>b</b>) assessing the performance of the SRC.</p>
Full article ">Figure 7
<p>Monthly average predicted sediment discharge using the developed SRC.</p>
Full article ">Figure 8
<p>Temporal variability of rainfall (R), streamflow (Q), and sediment discharge (Qs) in Dry Creek (2001–2022).</p>
Full article ">
9 pages, 5647 KiB  
Article
Discovery of Intrinsic Ferromagnetism Induced by Memory Effects in Low-Dimensional System
by Shaolong Zeng, Xuejin Wan, Yangfan Hu, Shijing Tan and Biao Wang
Fractal Fract. 2024, 8(11), 668; https://doi.org/10.3390/fractalfract8110668 (registering DOI) - 16 Nov 2024
Viewed by 183
Abstract
The impact of dynamic processes on equilibrium properties is a fundamental issue in condensed matter physics. This study investigates the intrinsic ferromagnetism generated by memory effects in the low-dimensional continuous symmetry Landau–Ginzburg model, demonstrating how memory effects can suppress fluctuations and stabilize long-range [...] Read more.
The impact of dynamic processes on equilibrium properties is a fundamental issue in condensed matter physics. This study investigates the intrinsic ferromagnetism generated by memory effects in the low-dimensional continuous symmetry Landau–Ginzburg model, demonstrating how memory effects can suppress fluctuations and stabilize long-range magnetic order. Our results provide compelling evidence that tuning dynamical processes can significantly alter the behavior of systems in equilibrium. We quantitatively evaluate how the emergence of the ferromagnetic phase depends on memory effects and confirm the presence of ferromagnetism through simulations of hysteresis loops, spontaneous magnetization, and magnetic domain structures in the 1D continuous symmetry Landau–Ginzburg model. This research offers both theoretical and numerical insights for identifying new phases of matter by dynamically modifying equilibrium properties. Full article
Show Figures

Figure 1

Figure 1
<p>Magnetic hysteresis loop at <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mo>−</mo> <mn>5</mn> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> (<b>b</b>) in the 1D continuous symmetry Landau–Ginzburg model for fractional temporal derivatives.</p>
Full article ">Figure 2
<p>Absolute value of magnetization in the z-direction <math display="inline"><semantics> <mrow> <mrow> <mo>|</mo> </mrow> <msub> <mi>M</mi> <mi>z</mi> </msub> <mrow> <mo>|</mo> </mrow> </mrow> </semantics></math> versus external field in the z-direction <math display="inline"><semantics> <msub> <mi>h</mi> <mi>z</mi> </msub> </semantics></math> for the 1D continuous symmetry Landau–Ginzburg model with fractional order <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>. The green dashed line corresponds to the position at 162,500 steps.</p>
Full article ">Figure 3
<p>Magnetic structure of the 1D continuous symmetry Landau–Ginzburg model in <a href="#fractalfract-08-00668-f002" class="html-fig">Figure 2</a> under an external field <math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mi>z</mi> </msub> <mo>=</mo> <mo>−</mo> <mn>0.1</mn> </mrow> </semantics></math> at step = 162,500. The color scale represents the projection of the order parameter in the z-direction.</p>
Full article ">Figure 4
<p>Spontaneous magnetization in the 1D continuous symmetry Landau–Ginzburg model with fractional temporal derivatives. Results averaged over 100 samples.</p>
Full article ">Figure 5
<p>The magnetic structure with <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.4</mn> </mrow> </semantics></math> (<b>a</b>) and <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.7</mn> </mrow> </semantics></math> (<b>b</b>). The arrows represent the order parameter directions at different positions, and the color scale represents the projection of the order parameter in the z-direction.</p>
Full article ">
9 pages, 11998 KiB  
Article
A Progressive Loss Decomposition Method for Low-Frequency Shielding of Soft Magnetic Materials
by Airu Ji and Jinji Sun
Materials 2024, 17(22), 5584; https://doi.org/10.3390/ma17225584 - 15 Nov 2024
Viewed by 173
Abstract
Energy loss in shielding soft magnetic materials at low frequencies (1–100 Hz) can cause fluctuations in the material’s magnetic field, and the resulting magnetic noise can interfere with the measurement accuracy and basic precision physics of biomagnetic signals. This places higher demands on [...] Read more.
Energy loss in shielding soft magnetic materials at low frequencies (1–100 Hz) can cause fluctuations in the material’s magnetic field, and the resulting magnetic noise can interfere with the measurement accuracy and basic precision physics of biomagnetic signals. This places higher demands on the credibility and accuracy of loss separation predictions. The current statistical loss theory (STL) method tends to ignore the high impact of the excitation dependence of quasi-static loss in the low-frequency band on the prediction accuracy. STL simultaneously fits and predicts multiple unknown quantities, causing its results to occasionally fall into the value boundary, and the credibility is low in the low-frequency band and with less data. This paper proposes a progressive loss decomposition (PLD) method. Through multi-step progressive predictions, the hysteresis loss simulation coefficients are first determined. The experimental data of the test ring verifies the credibility of PLD’s prediction of the two hysteresis coefficients, improving the inapplicability of the STL method. In addition, we use the proposed method to obtain the prediction results of the low-frequency characteristics of the loss of a variety of typical soft magnetic materials, providing a reference for analyzing the loss characteristics of materials. Full article
Show Figures

Figure 1

Figure 1
<p>Test instrument and magnetic ring size.</p>
Full article ">Figure 2
<p>Loss separation process. (<b>a</b>) Original test data (including error bars). (<b>b</b>) First step fitting. (<b>c</b>) Second step fitting: intercept of each curve in (<b>b</b>). The color of symbol ’x’ corresponds to the legend in (<b>b</b>). (<b>d</b>) Total loss fitting surface: the fitting surface uses parula colormap to represent the size of the predicted value. (<b>e</b>) Loss separation prediction results.</p>
Full article ">Figure 3
<p>Loss separation prediction steps for Co-based amorphous. (The symbol ‘x’ represents the intercept of each curve. The fitting surface represents the size of the predicted value using parula colormap).</p>
Full article ">Figure 4
<p>Loss separation prediction steps for Finemet NANO. (The symbol ‘x’ represents the intercept of each curve. The fitting surface represents the size of the predicted value using parula colormap).</p>
Full article ">Figure 5
<p>Loss separation prediction steps for ferrtie. (The symbol ‘x’ represents the intercept of each curve. The fitting surface represents the size of the predicted value using parula colormap).</p>
Full article ">Figure 6
<p>Separation of power loss components of 4 materials.</p>
Full article ">
16 pages, 10754 KiB  
Article
Unveiling Frequency-Dependent Electromechanical Dynamics in Ferroelectric BaTiO3 Nanofilm with a Core-Shell Structure
by Mingran Zhang, Rui Ma, Jianqiang Zhou, Yuanxiang Zhang, Jie Wang and Shengbin Weng
Coatings 2024, 14(11), 1437; https://doi.org/10.3390/coatings14111437 - 12 Nov 2024
Viewed by 316
Abstract
Diverse domain patterns significantly influence the nonlinear electromechanical behaviors of ferroelectric nanomaterials, with polarization switching under strong electric fields being inherently a frequency-dependent phenomenon. Nevertheless, research in this area remains limited. In this study, we present a phase-field investigation of frequency-dependent electromechanical dynamics [...] Read more.
Diverse domain patterns significantly influence the nonlinear electromechanical behaviors of ferroelectric nanomaterials, with polarization switching under strong electric fields being inherently a frequency-dependent phenomenon. Nevertheless, research in this area remains limited. In this study, we present a phase-field investigation of frequency-dependent electromechanical dynamics of a polycrystalline BaTiO3 nanofilm with a core-shell structure, subjected to applied frequencies ranging from 1 to 80 kHz. Our findings elucidate the microstructural mechanisms underlying the electromechanical behaviors observed in these materials. The effect of the grain size and the strains effect are also taken into account. Hysteresis and butterfly loops exhibit a marked change in shape as the frequency changes. We discuss the underlying domain-switching dynamics as a basis for evaluating such frequency-dependent properties. In addition, we examine the scaling behaviors of the dynamic hysteresis and the influence of grain boundaries on the domain structure. We can also observe from hysteresis loops that the remnant polarization and coercive field significantly diminish when grain sizes decrease from 60 to 5 nm. A smaller grain size of the nanofilm yields a larger percentage of the dielectric grain boundary, which “dilutes” the overall ferroelectricity of the film. A vortex domain structure is more likely to form at low frequency and a small grain size. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of a the BaTiO<sub>3</sub> polycrystalline nano film with 22 grains (core-shell) and marked orientation of each grain mentioned.</p>
Full article ">Figure 2
<p>The calculated (<b>a</b>) hysteresis loops and (<b>b</b>) butterfly loops at different frequencies of the BaTiO<sub>3</sub> nano-film under an in-plane strain of −0.1%.</p>
Full article ">Figure 3
<p>The calculated (<b>a</b>) hysteresis loops and (<b>b</b>) butterfly loops at different frequencies of the BaTiO<sub>3</sub> nano-film under an in-plane strain of −2.23%.</p>
Full article ">Figure 4
<p>The frequency-related parameters of (<b>a</b>) the coercive field and remnant polarization; (<b>b</b>) the piezoelectric coefficient and dielectric constant. (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>κ</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>d</mi> </mrow> <mrow> <mn>33</mn> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mrow> </mrow> </semantics></math>).</p>
Full article ">Figure 5
<p>Scaling plots of loop area <math display="inline"><semantics> <mrow> <mi>A</mi> </mrow> </semantics></math> against <math display="inline"><semantics> <mrow> <mi>f</mi> </mrow> </semantics></math> for BaTiO<sub>3</sub> polycrystalline nanofilm at low frequency and high frequency.</p>
Full article ">Figure 6
<p>(<b>a</b>) The hysteresis loops and (<b>b</b>) the butterfly loops with different grain sizes of polycrystalline nanofilms.</p>
Full article ">Figure 7
<p>The calculated properties against the grain size: (<b>a</b>) the coercive field, the dielectric permittivity, and piezoelectric coefficient at the zero electric field; (<b>b</b>) the remnant polarization and other experimental data (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>κ</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mrow> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>d</mi> </mrow> <mrow> <mn>33</mn> </mrow> <mrow> <mn>0</mn> </mrow> </msubsup> <mo>=</mo> <mrow> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> <mo>/</mo> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mrow> </mrow> </semantics></math>) [<a href="#B42-coatings-14-01437" class="html-bibr">42</a>,<a href="#B48-coatings-14-01437" class="html-bibr">48</a>,<a href="#B49-coatings-14-01437" class="html-bibr">49</a>,<a href="#B50-coatings-14-01437" class="html-bibr">50</a>,<a href="#B51-coatings-14-01437" class="html-bibr">51</a>].</p>
Full article ">Figure 8
<p>The microstructure of the BaTiO<sub>3</sub> nanofilm at frequency 0.1 kHz (<b>a</b>) <span class="html-italic">E</span> = 0; (<b>b</b>) <span class="html-italic">E</span> = −0.25<span class="html-italic">E</span><sub>0</sub>; (<b>c</b>) <span class="html-italic">E</span> = −8<span class="html-italic">E</span><sub>0</sub>; (<b>d</b>) the corresponding positions of different states in the hysteresis loop.</p>
Full article ">Figure 9
<p>The microstructure of the BaTiO<sub>3</sub> nanofilm at frequency 50 kHz (<b>a</b>) <span class="html-italic">E</span> = 0; (<b>b</b>) <span class="html-italic">E</span> = −7.8<span class="html-italic">E</span><sub>0</sub>; (<b>c</b>) <span class="html-italic">E</span> = 0; (<b>d</b>) the corresponding positions of different states in the hysteresis loop.</p>
Full article ">Figure 10
<p>The remnant-state polarization distribution of the polycrystalline nanofilm with the grain sizes being (<b>a</b>) 10 nm; (<b>b</b>) 15 nm; (<b>c</b>) 30 nm.</p>
Full article ">Figure 11
<p>The microstructure of the nanofilm at frequency 5 kHz under in-plane strains −2.23%: (<b>a</b>) the remnant-state polarization; (<b>b</b>) the coercive field state.</p>
Full article ">
17 pages, 37894 KiB  
Article
High-Precision Rotor Position Fitting Method of Permanent Magnet Synchronous Machine Based on Hall-Effect Sensors
by Kaining Qu, Pengfei Pang and Wei Hua
Energies 2024, 17(22), 5625; https://doi.org/10.3390/en17225625 - 10 Nov 2024
Viewed by 499
Abstract
The high-performance vector control technology of permanent magnet synchronous machines (PMSMs) relies on high-precision rotor position. The Hall-effect sensor has the advantages of low cost, simple installation, and strong anti-interference ability. However, it can only provide six discrete rotor angles in an electrical [...] Read more.
The high-performance vector control technology of permanent magnet synchronous machines (PMSMs) relies on high-precision rotor position. The Hall-effect sensor has the advantages of low cost, simple installation, and strong anti-interference ability. However, it can only provide six discrete rotor angles in an electrical cycle, which makes high-precision vector control of PMSMs difficult. Hence, to obtain the necessary rotor position of PMSMs, a rotor position fitting method combining the Hall signal and machine flux information is proposed. Firstly, the rotor position signal output by the Hall-effect sensors is used to calibrate and update the stator flux obtained under pure integration. Then, based on the corrected stator flux and its relationship with current and angle, the rotor position and speed are obtained. Experimental verification shows that the rotor position observer combining Hall signal and flux information can reduce the initial value bias and integral drift caused by traditional average speed method hysteresis and pure integration method calculation of flux and can quickly and accurately track and estimate the rotor position, achieving high-performance vector control of PMSMs. Full article
(This article belongs to the Special Issue Designs and Control of Electrical Machines and Drives)
Show Figures

Figure 1

Figure 1
<p>Hall-effect Sensors.</p>
Full article ">Figure 2
<p>Hall installation mode and output signals (<b>a</b>) 120° Hall-effect installation, (<b>b</b>) Hall-effect output signals.</p>
Full article ">Figure 3
<p>Relation of Hall signal and rotor position.</p>
Full article ">Figure 4
<p>Estimation principle by average velocity method.</p>
Full article ">Figure 5
<p>Hall signal delay under digital control system.</p>
Full article ">Figure 6
<p>Structure of position vector tracking observer.</p>
Full article ">Figure 7
<p>Bode diagram of position vector tracking observer based on back EMF.</p>
Full article ">Figure 8
<p>Stator flux observer combined with flux linkage information and Hall signal.</p>
Full article ">Figure 9
<p>Input voltage calculation for the flux observer.</p>
Full article ">Figure 10
<p>Rotor position observer combined with flux linkage information and Hall signal.</p>
Full article ">Figure 11
<p>The experiment platform.</p>
Full article ">Figure 12
<p>Flux linkage waveforms under three conditions (<b>a</b>) Flux linkage before correction, (<b>b</b>) Flux linkage at discrete Hall points, (<b>c</b>) Flux linkage under Hall signal correction.</p>
Full article ">Figure 13
<p>Comparison of no-load starting rotor positions (<b>a</b>) Average speed, (<b>b</b>) Vector tracking, (<b>c</b>) Flux-Hall.</p>
Full article ">Figure 14
<p>Comparison of on-rated-load starting rotor positions.</p>
Full article ">Figure 15
<p>Estimated rotor angle errors under three initial positions (<b>a</b>) The Hall intermediate angle, (<b>b</b>) 25° advanced, (<b>c</b>) 25° delayed.</p>
Full article ">Figure 16
<p>Comparison of <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>q</mi> </mrow> </semantics></math>-frame currents waveforms. Comparison of <math display="inline"><semantics> <mrow> <mi>d</mi> <mi>q</mi> </mrow> </semantics></math>-frame currents waveforms (<b>a</b>) Average speed, (<b>b</b>) Flux-Hall.</p>
Full article ">Figure 17
<p>Comparison of rotor positions under a speed change from 300/min to 1800/min (<b>a</b>) Average speed, (<b>b</b>) Flux-Hall, (<b>c</b>) Angle error.</p>
Full article ">Figure 18
<p>Rotor position during forward and reverse switching (<b>a</b>) Average speed, (<b>b</b>) Flux-Hall.</p>
Full article ">Figure 19
<p>Stator current and rotor position at 750/min (<b>a</b>) Average speed, (<b>b</b>,<b>c</b>) Flux-Hall.</p>
Full article ">Figure 20
<p>Rotor position under different inductance (<b>a</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>L</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mn>0.8</mn> <mi>L</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>L</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mn>1.2</mn> <mi>L</mi> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>L</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mi>L</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 21
<p>Rotor position under different resistance (<b>a</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>R</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mn>0.8</mn> <mi>R</mi> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>R</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mn>1.2</mn> <mi>R</mi> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>R</mi> </mrow> <mo stretchy="false">^</mo> </mover> <mo>=</mo> <mi>R</mi> </mrow> </semantics></math>.</p>
Full article ">
10 pages, 830 KiB  
Article
Coexistence of Superconductivity and Magnetic Ordering in the In–Ag Alloy Under Nanoconfinement
by Marina V. Likholetova, Elena V. Charnaya, Evgenii V. Shevchenko, Yurii A. Kumzerov and Aleksandr V. Fokin
Nanomaterials 2024, 14(22), 1792; https://doi.org/10.3390/nano14221792 - 7 Nov 2024
Viewed by 501
Abstract
The impact of the interface phenomena on the properties of nanostructured materials is the focus of modern physics. We studied the magnetic properties of the nanostructured In–Ag alloy confined within a porous glass. The alloy composition was close to the eutectic point in [...] Read more.
The impact of the interface phenomena on the properties of nanostructured materials is the focus of modern physics. We studied the magnetic properties of the nanostructured In–Ag alloy confined within a porous glass. The alloy composition was close to the eutectic point in the indium-rich range of the phase diagram. Temperature dependences of DC magnetization evidenced two superconducting transitions at 4.05 and 3.38 K. The magnetization isotherms demonstrated the superposition of two hysteresis loops with low and high critical fields below the second transition, a single hysteresis between the transitions and ferromagnetism with weak remanence in the normal state of the alloy. The shape of the loop seen below the second transition, which closes at a low magnetic field, corresponded to the intermediate state of the type-I superconductor. It was ascribed to strongly linked indium segregates. The loop observed below the first transition is referred to as type-II superconductivity. The secondary and tertiary magnetization branches measured at decreasing and increasing fields were shifted relative to each other, revealing the proximity of superconducting and ferromagnetic phases at the nanometer scale. This phenomenon was observed for the first time in the alloy, whose components were not magnetic in bulk. The sign of the shift shows the dominant role of the stray fields of ferromagnetic regions. Ferromagnetism was suggested to emerge at the interface between the In and AgIn2 segregates. Full article
(This article belongs to the Section Nanocomposite Materials)
Show Figures

Figure 1

Figure 1
<p>Pore size distribution in the porous glass.</p>
Full article ">Figure 2
<p>Temperature dependences of susceptibilities measured under the ZFC (black symbols and lines), FCC (blue symbols and lines), and FCW (red symbols and lines) protocols at fields of 10 (<b>a</b>), 50 (<b>b</b>), and 100 (<b>c</b>) Oe.</p>
Full article ">Figure 3
<p>Temperature dependences of susceptibilities measured at 300 and 500 Oe under the ZFC (black symbols and lines), FCC (blue symbols and lines), and FCW (red symbols and lines) protocols.</p>
Full article ">Figure 4
<p>Central parts of isotherms of magnetizations obtained at temperatures 1.8 (<b>a</b>), 3.3 (<b>b</b>), 3.6 (<b>c</b>), and 8 (<b>d</b>) K. The arrows indicate the directions of ramping the field. The red, green, and blue symbols and lines correspond to the virgin, secondary, and tertiary magnetizations, respectively. The insets present the magnetization curves on a larger scale.</p>
Full article ">Figure 5
<p>X-ray pattern of the porous glass/In–Ag alloy nanocomposite.</p>
Full article ">Figure 6
<p>The separated central hysteresis loops for temperatures 1.8 (<b>a</b>) and 3.3 (<b>b</b>) K. The arrows indicate the directions of ramping the field. The red, green, and blue symbols and lines correspond to the virgin, secondary, and tertiary magnetizations, respectively.</p>
Full article ">
15 pages, 2839 KiB  
Article
Computational Modeling of U-Shaped Seismic Dampers for Structural Damage Mitigation
by Víctor Tuninetti, Álvaro Gómez, Flavia Bustos, Angelo Oñate, Jorge Hinojosa, Calogero Gallo, Anne-Marie Habraken and Laurent Duchêne
Appl. Sci. 2024, 14(22), 10238; https://doi.org/10.3390/app142210238 - 7 Nov 2024
Viewed by 433
Abstract
U-shaped seismic dampers, passive metallic devices that dissipate energy by cyclic plastic deformation, are designed to mitigate the effects of seismic loads on structures. This study focuses on the development of an advanced computational model of a U-shaped damper, chosen for its unique [...] Read more.
U-shaped seismic dampers, passive metallic devices that dissipate energy by cyclic plastic deformation, are designed to mitigate the effects of seismic loads on structures. This study focuses on the development of an advanced computational model of a U-shaped damper, chosen for its unique design of variable thickness and width, which contributes to its superior performance. The simulation uses nonlinear finite element analysis and a bilinear hardening model calibrated to the actual stress–strain curve of the low-carbon steel. To ensure accuracy, a rigorous mesh convergence analysis is performed to quantify numerical prediction errors and establish a model suitable for predicting local deformation phenomena, including strain and stress fields, throughout the displacement-based loading protocol. Mesh sensitivity analysis, performed by examining the equivalent stress and cumulative plastic strain, derives the damper hysteresis curve and confirms the convergence criteria of the mesh within the experimentally observed plastic response range of the material. The resulting computational model is a novel contribution that provides reliable predictions of local inhomogeneous deformation and energy dissipation, essential for optimizing damper design and performance through more sophisticated damage-fatigue models that guarantee the lifetime of a damper. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) UD-40 model with boundary conditions applied. (<b>b</b>) Upper and front sight of UD-40 USSD with dimensions in mm. (<b>c</b>) Load protocol applied to the damper. (<b>d</b>) True stress–strain curve from experimental data reported for JIS-SM490 fitted with the bilinear hardening model.</p>
Full article ">Figure 2
<p>Finite element models of the U-shaped damper with increasing mesh density through the thickness.</p>
Full article ">Figure 3
<p>Maximum equivalent von Mises stress for the different mesh sizes in convergence analysis.</p>
Full article ">Figure 4
<p>Mesh convergence analysis of (<b>a</b>) estimated fractional error (E1) vs. element size for 1 cycle, (<b>b</b>) 5 cycles, and (<b>c</b>) 10 cycles. (<b>d</b>) Maximum equivalent stress, (<b>e</b>) AEPS, (<b>f</b>) equivalent plastic strain, and (<b>g</b>) strain energy vs. number of elements in the damper thickness.</p>
Full article ">Figure 5
<p>Fields of (<b>a</b>) equivalent von Mises stress, (<b>b</b>) equivalent cumulative plastic strain (AEPS), (<b>c</b>) equivalent plastic strain, and (<b>d</b>) strain energy on the ultimate applied displacement.</p>
Full article ">Figure 6
<p>(<b>a</b>) Hysteresis curve and (<b>b</b>) plastic strain energy vs. displacement.</p>
Full article ">
7 pages, 1951 KiB  
Proceeding Paper
Development of Food Hydrogels with Andean Purple Corn (Zea mays L.) Extracts and Cushuro (Nostoc sphaericum) Polysaccharide: Rheological Characterization
by Cecilia A. Arenas and Nancy A. Chasquibol
Biol. Life Sci. Forum 2024, 37(1), 1; https://doi.org/10.3390/blsf2024037001 - 29 Oct 2024
Viewed by 279
Abstract
Andean purple corn (Zea mays L.) is an ancient native Peruvian crop that is currently used in Peruvian gastronomy. Cushuro is a cyanobacteria from the Andean lakes of Peru. They have considerable amounts of bioactive compounds that can improve the physicochemical properties [...] Read more.
Andean purple corn (Zea mays L.) is an ancient native Peruvian crop that is currently used in Peruvian gastronomy. Cushuro is a cyanobacteria from the Andean lakes of Peru. They have considerable amounts of bioactive compounds that can improve the physicochemical properties of foods. The objective of this research was to characterize the rheological and functional properties of food hydrogels developed with purple corn extracts, red prickly pear fruit pulp, and cushuro polysaccharide (CP). Acid-soluble polysaccharides obtained from the Nostoc sphaericum variety from Ancash, Peru, as well as Peruvian purple corn extracts, were used. Food hydrogels at concentrations ranging from 0.5% to 3.5% (w/v) were elaborated by dispersing the polysaccharides in a 4:1 extract/pulp (v/v) ratio. Likewise, control samples with Tara (Caesalpinia spinosa) gum (TG) were made. The effect of hydrocolloid concentration (0.5; 1.5; 2.5; 3.5%) on the rheological properties was evaluated using a unifactorial design. CP and TG hydrogels exhibited a shear-thinning nature, a concentration-dependent yield point (0.02–29.91 Pa; 2.01–508.39 Pa), and high antioxidant activity and phenolic content. Adding CP revealed slow structural regeneration, while TG showed thixotropy and a symmetric hysteresis loop. CP gels showed a fluid-like structure with viscoelastic properties (G″ > G′) even in the highest concentration evaluated (3.5%), contrary to TG gel that had a more solid (gel-like) structure (G′ > G″) at a low concentration (1.5%). These results showed a suitable rheological profile and desirable properties of the food hydrogels development for the functional food industry and processing. Full article
Show Figures

Figure 1

Figure 1
<p>Flow (<b>A</b>) and viscosity (<b>B</b>) curves of CP3.5 and TG1.5 hydrogels. CP: cushuro polysaccharide; TG: Tara gum.</p>
Full article ">Figure 2
<p>Thixotropy curves (<b>A</b>) and hysteresis loop (<b>B</b>) of CP3.5 and TG1.5 hydrogels. CP: cushuro polysaccharide; TG: Tara gum.</p>
Full article ">Figure 3
<p>Amplitude sweeps of CP3.5 and TG1.5 hydrogels.</p>
Full article ">Figure 4
<p>Frequency sweeps (<b>A</b>) and complex viscosity (<b>B</b>) curves of CP3.5 and TG1.5 hydrogels. CP: cushuro polysaccharide; TG: Tara gum.</p>
Full article ">
19 pages, 2406 KiB  
Article
FPGA Realization of a Fractional-Order Model of Universal Memory Elements
by Opeyemi-Micheal Afolabi, Vincent-Ademola Adeyemi, Esteban Tlelo-Cuautle and Jose-Cruz Nuñez-Perez
Fractal Fract. 2024, 8(10), 605; https://doi.org/10.3390/fractalfract8100605 - 18 Oct 2024
Viewed by 1193
Abstract
This paper addresses critical gaps in the digital implementations of fractional-order memelement emulators, particularly given the challenges associated with the development of solid-state devices using nanomaterials. Despite the potentials of these devices for industrial applications, the digital implementation of fractional-order models has received [...] Read more.
This paper addresses critical gaps in the digital implementations of fractional-order memelement emulators, particularly given the challenges associated with the development of solid-state devices using nanomaterials. Despite the potentials of these devices for industrial applications, the digital implementation of fractional-order models has received limited attention. This research contributes to bridging this knowledge gap by presenting the FPGA realization of the memelements based on a universal voltage-controlled circuit topology. The digital emulators successfully exhibit the pinched hysteresis behaviors of memristors, memcapacitors, and meminductors, showing the retention of historical states of their constitutive electronic variables. Additionally, we analyze the impact of the fractional-order parameters and excitation frequencies on the behaviors of the memelements. The design methodology involves using Xilinx System Generator for DSP blocks to lay out the architectures of the emulators, with synthesis and gate-level implementation performed on the Xilinx Artix-7 AC701 Evaluation kit, where resource utilization on hardware accounts for about 1% of available hardware resources. Further hardware analysis shows successful timing validation and low power consumption across all designs, with an average on-chip power of 0.23 Watts and average worst negative slack of 0.6 ns against a 5 ns constraint. We validate these results with Matlab 2020b simulations, which aligns with the hardware models. Full article
(This article belongs to the Section Engineering)
Show Figures

Figure 1

Figure 1
<p>Memory effect or time non-locality. Source: own elaboration.</p>
Full article ">Figure 2
<p>Square symmetric diagram of all-known fundamental electronic elements. Source: own elaboration.</p>
Full article ">Figure 3
<p>Digital FOMR: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the fractional integral of the input voltage <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> </msub> </semantics></math> configuration in its steady-state response, (<b>c</b>) the hardware co-simulation block, and (<b>d</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 3 Cont.
<p>Digital FOMR: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the fractional integral of the input voltage <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>A</mi> <mi>B</mi> </mrow> </msub> </semantics></math> configuration in its steady-state response, (<b>c</b>) the hardware co-simulation block, and (<b>d</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 4
<p>Digital FOMC: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the hardware co-simulation block, and (<b>c</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 4 Cont.
<p>Digital FOMC: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the hardware co-simulation block, and (<b>c</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 5
<p>Digital FOMI: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the digital integer-order integrator, (<b>c</b>) the hardware co-simulation block, and (<b>d</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 5 Cont.
<p>Digital FOMI: (<b>a</b>) the complete design of the emulator, (<b>b</b>) the digital integer-order integrator, (<b>c</b>) the hardware co-simulation block, and (<b>d</b>) an evaluation with a voltage source. Source: own elaboration.</p>
Full article ">Figure 6
<p>Dynamics of <math display="inline"><semantics> <mrow> <msup> <mi>J</mi> <mi>α</mi> </msup> <mrow> <mo>(</mo> <mi>A</mi> <msup> <mi>ω</mi> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> <mi>s</mi> <mi>i</mi> <mi>n</mi> <mrow> <mo>(</mo> <mi>ω</mi> <mi>t</mi> <mo>)</mo> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> for <span class="html-italic">f</span> = 7 kHz, <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.95, <span class="html-italic">A</span> = 1.5 V, and step size = <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>−</mo> <mi>i</mi> </mrow> </semantics></math> characteristics of FOMR for different excitation frequencies of <span class="html-italic">f</span> = 0.4 kHz, 0.8 kHz, and 1.2 kHz when <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.95. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">Figure 8
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi>v</mi> <mo>−</mo> <mi>i</mi> </mrow> </semantics></math> characteristics of FOMR for different order values <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.90, 0.95, and 1 when <span class="html-italic">f</span> = 0.4 kHz. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">Figure 9
<p><math display="inline"><semantics> <mrow> <mi>q</mi> <mo>−</mo> <mi>v</mi> </mrow> </semantics></math> characteristics of FOMC for different excitation frequencies of <span class="html-italic">f</span> = 1.5 kHz, 2 kHz, and 2.5 kHz when <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.95. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">Figure 10
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi>q</mi> <mo>−</mo> <mi>v</mi> </mrow> </semantics></math> characteristics of FOMC for different order values <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.90, 0.95, and 1 when <span class="html-italic">f</span> = 2.5 kHz. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">Figure 11
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>−</mo> <mi>i</mi> </mrow> </semantics></math> characteristics of FOMI for different excitation frequencies of <span class="html-italic">f</span> = 5 kHz, 6 kHz, and 7 kHz when <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.95. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">Figure 12
<p>(<b>a</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> <mo>−</mo> <mi>i</mi> </mrow> </semantics></math> characteristics of FOMI for different order values <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.90, 0.95, and 1 when <span class="html-italic">f</span> = 5 kHz. (<b>a</b>) Matlab. (<b>b</b>) Hardware Co-simulation. Source: own elaboration.</p>
Full article ">
23 pages, 5832 KiB  
Article
Usage of Machine Learning Techniques to Classify and Predict the Performance of Force Sensing Resistors
by Angela Peña, Edwin L. Alvarez, Diana M. Ayala Valderrama, Carlos Palacio, Yosmely Bermudez and Leonel Paredes-Madrid
Sensors 2024, 24(20), 6592; https://doi.org/10.3390/s24206592 - 13 Oct 2024
Viewed by 959
Abstract
Recently, there has been a huge increase in the different ways to manufacture polymer-based sensors. Methods like additive manufacturing, microfluidic preparation, and brush painting are just a few examples of new approaches designed to improve sensor features like self-healing, higher sensitivity, reduced drift [...] Read more.
Recently, there has been a huge increase in the different ways to manufacture polymer-based sensors. Methods like additive manufacturing, microfluidic preparation, and brush painting are just a few examples of new approaches designed to improve sensor features like self-healing, higher sensitivity, reduced drift over time, and lower hysteresis. That being said, we believe there is still a lot of potential to boost the performance of current sensors by applying modeling, classification, and machine learning techniques. With this approach, final sensor users may benefit from inexpensive computational methods instead of dealing with the already mentioned manufacturing routes. In this study, a total of 96 specimens of two commercial brands of Force Sensing Resistors (FSRs) were characterized under the error metrics of drift and hysteresis; the characterization was performed at multiple input voltages in a tailored test bench. It was found that the output voltage at null force (Vo_null) of a given specimen is inversely correlated with its drift error, and, consequently, it is possible to predict the sensor’s performance by performing inexpensive electrical measurements on the sensor before deploying it to the final application. Hysteresis error was also studied in regard to Vo_null readings; nonetheless, a relationship between Vo_null and hysteresis was not found. However, a classification rule base on k-means clustering method was implemented; the clustering allowed us to distinguish in advance between sensors with high and low hysteresis by relying solely on Vo_null readings; the method was successfully implemented on Peratech SP200 sensors, but it could be applied to Interlink FSR402 sensors. With the aim of providing a comprehensive insight of the experimental data, the theoretical foundations of FSRs are also presented and correlated with the introduced modeling/classification techniques. Full article
(This article belongs to the Special Issue Advanced Flexible Electronics for Sensing Application)
Show Figures

Figure 1

Figure 1
<p>Sketch of an unloaded (<b>left</b>) and a loaded (<b>right</b>) Force Sensing Resistor. An external applied force (<span class="html-italic">F</span>) causes a reduction in the interparticle separation from <span class="html-italic">s</span> down to <span class="html-italic">s</span>-Δ<span class="html-italic">s</span>; this applies to all the tunneling paths along the nanocomposite. The constriction resistance spots are signaled as gray arrow marks. The constriction resistance is also modified by the applied force.</p>
Full article ">Figure 2
<p>Sketches of FSRs with two different sensing mechanisms. (<b>a</b>) Quantum tunneling dominates as agglomerated particles are separated from each other, but connected through multiple tunneling paths. (<b>b</b>) Percolation dominates as particles form connection bridged between electrodes; there are only a few tunneling paths along the nanocomposite.</p>
Full article ">Figure 3
<p>Sketch of common errors in FSRs measured from sensor conductivity (<span class="html-italic">σ</span>). (<b>a</b>) Drift error occurring after one hour of constant loading. (<b>b</b>) Hysteresis error occurring during loading (solid line) and unloading (dashed line) stages.</p>
Full article ">Figure 4
<p>Photographs of the mechanical setup. (<b>a</b>) Overview of the testbench; (<b>b</b>) Zoom-in photo depicting the linear motor for applying forces to the bunch of sensors. (<b>c</b>) Photograph of FSRs installed inside sensor holders, Peratech SP200 (yellow) and Interlink FSR402 (orange). (<b>d</b>) Zoom-in photo depicting the sensors inside the chamber and the spring for mechanical compliance. (<b>e</b>) Photograph of two side-by-side sensor holders showing the puck (top side) and the notch (bottom side). (<b>f</b>) Custom design element for holding aligned the sensor holders, and the spring. A FSR was placed near the element for comparison purposes.</p>
Full article ">Figure 5
<p>Inverting amplifier for measuring the conductivity of 16 FSRs. The ADG444 was added in order to handle the sensors in a time-multiplexed fashion.</p>
Full article ">Figure 6
<p>Drift error (<span class="html-italic">d.e.</span>) at multiple input voltages for the Interlink FSR402 (<b>a</b>–<b>c</b>) and QTC Peratech SP200 (<b>d</b>–<b>f</b>) sensors.</p>
Full article ">Figure 7
<p>Drift error (<span class="html-italic">d.e.</span>) measured at <span class="html-italic">V<sub>i</sub></span> = 5 V with three superimposed trendlines. (<b>a</b>) Interlink FSR402, (<b>b</b>) QTC Peratech SP200.</p>
Full article ">Figure 8
<p>Flowchart summarizing the process for measuring and modeling the drift error in FSRs.</p>
Full article ">Figure 9
<p>Hysteresis error (<span class="html-italic">h.e.</span>) measured at multiple input voltage for Interlink FSR402 (<b>a</b>–<b>c</b>) and QTC Peratech SP200 (<b>d</b>–<b>f</b>) sensors.</p>
Full article ">Figure 10
<p>Plots of two Interlink sensors showing the relationship between <span class="html-italic">V<sub>o_null</sub></span> and <span class="html-italic">V<sub>i</sub></span> at null applied force. (<b>a</b>) Parabolic behavior with <span class="html-italic">g</span>/<span class="html-italic">f</span> =0.44. (<b>b</b>) Linear behavior with <span class="html-italic">g</span>/<span class="html-italic">f</span> = 8.78.</p>
Full article ">Figure 11
<p>Elbow method results for the <span class="html-italic">h.e.</span> data of Peratech sensors at <span class="html-italic">V<sub>i</sub></span> = 5 V.</p>
Full article ">Figure 12
<p>Result of <span class="html-italic">k</span>-means clustering method using <span class="html-italic">k</span> = 2 for Peratech SP200 sensors at <span class="html-italic">V<sub>i</sub></span> = 5 V.</p>
Full article ">Figure 13
<p>Result of <span class="html-italic">k</span>-means clustering method using <span class="html-italic">k</span> = 3 for Peratech SP200 sensors at <span class="html-italic">V<sub>i</sub></span> = 5 V.</p>
Full article ">Figure 14
<p>Result of <span class="html-italic">k</span>-means clustering method using <span class="html-italic">k</span> = 4 for Peratech SP200 sensors at <span class="html-italic">V<sub>i</sub></span> = 5 V.</p>
Full article ">Figure 15
<p>Flowchart summarizing the process for measuring and assessing the hysteresis error in Peratech SP200 sensors.</p>
Full article ">
11 pages, 6060 KiB  
Article
Investigation of Asymmetric Flow of a Slender Body with Low-Aspect Ratio Fins Having Large Deflection Angles
by Yonghong Li, Lin Zhang, Chuan Gao, Jilong Zhu and Bin Dong
Aerospace 2024, 11(10), 835; https://doi.org/10.3390/aerospace11100835 - 10 Oct 2024
Viewed by 494
Abstract
To understand the asymmetric flow of a slender body with low-aspect ratio fins, a wind tunnel experiment was carried out, and the asymmetric flow was observed when the pair of fins had a symmetric deflection angle of 30° at a small angle of [...] Read more.
To understand the asymmetric flow of a slender body with low-aspect ratio fins, a wind tunnel experiment was carried out, and the asymmetric flow was observed when the pair of fins had a symmetric deflection angle of 30° at a small angle of attack and zero sideslip angle at transonic speeds. The unsteady characteristics of flow around the moving fins, especially for the evolution of the asymmetric flow, was carefully numerically investigated via the RANS method. To verify the numerical method, the experimental steady wind tunnel data of the NACA 0012 airfoil with sinusoidal pitching motion were adopted. A hysteresis loop exists as a function of the deflection angle during the upstroke and downstroke motions. The side force is periodic due to the asymmetric flow peaks at the downstroke and their peak value appeared at around δz = 25°, which was independent of the deflection frequency. As the deflection frequency increased, the asymmetric flow formed at a higher deflection angle during the upstroke motion, but decayed at a lower deflection angle during the downstroke motion, resulting in a more significant unsteady hysteresis effect. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>The sketch of the test model.</p>
Full article ">Figure 2
<p>The deflection of the pair of fins on the leeward and windward side with <span class="html-italic">δz</span> = 30°.</p>
Full article ">Figure 3
<p>The pitching moment and side force of the models with different pitching deflection angles against a range of angles of attack at <span class="html-italic">M</span> = 0.95, <span class="html-italic">β</span> = 0°. The uncertainties of C<sub>y</sub> and C<sub>m</sub> are 0.002 and 0.01, respectively.</p>
Full article ">Figure 4
<p>The side force coefficients under various Mach numbers at zero angle of attack, <span class="html-italic">β</span> = 0° (<span class="html-italic">δz</span> = 30°). The uncertainty of C<sub>y</sub> is 0.002.</p>
Full article ">Figure 5
<p>The original images of the PIV results at <span class="html-italic">β</span> = 0° (<span class="html-italic">δz</span> = 30°).</p>
Full article ">Figure 6
<p>The mesh topology of the fins and the rail body (2.4 million cells).</p>
Full article ">Figure 7
<p>Comparisons of the side forces between the experimental data and the CFD results of the model with <span class="html-italic">δz</span> = 30°.</p>
Full article ">Figure 8
<p>The stagnation pressure contours of type sections around the leeward and windward pair of fins.</p>
Full article ">Figure 9
<p>Mach number distributions of typical sections around the fins at <span class="html-italic">M</span> = 0.95.</p>
Full article ">Figure 10
<p>Comparisons of the experimental data with the present results.</p>
Full article ">Figure 11
<p>Evolution of the entropy flow field, <span class="html-italic">δz</span>(<span class="html-italic">t</span>) = 20° + 10° sin(2<span class="html-italic">πk∙t</span>), <span class="html-italic">k</span> = 0.044.</p>
Full article ">Figure 12
<p>Evolution of the side force coefficient, <span class="html-italic">δz</span>(<span class="html-italic">t</span>) = 20° + 10° sin(2<span class="html-italic">πk∙t</span>), <span class="html-italic">k</span> = 0.044.</p>
Full article ">Figure 13
<p>Time evolution of the side force coefficient of the fins with different deflection frequencies.</p>
Full article ">
13 pages, 3708 KiB  
Article
Nonlinear Modeling of a Piezoelectric Actuator-Driven High-Speed Atomic Force Microscope Scanner Using a Variant DenseNet-Type Neural Network
by Thi Thu Nguyen, Luke Oduor Otieno, Oyoo Michael Juma, Thi Ngoc Nguyen and Yong Joong Lee
Actuators 2024, 13(10), 391; https://doi.org/10.3390/act13100391 - 2 Oct 2024
Viewed by 491
Abstract
Piezoelectric actuators (PEAs) are extensively used for scanning and positioning in scanning probe microscopy (SPM) due to their high precision, simple construction, and fast response. However, there are significant challenges for instrument designers due to their nonlinear properties. Nonlinear properties make precise and [...] Read more.
Piezoelectric actuators (PEAs) are extensively used for scanning and positioning in scanning probe microscopy (SPM) due to their high precision, simple construction, and fast response. However, there are significant challenges for instrument designers due to their nonlinear properties. Nonlinear properties make precise and accurate control difficult in cases where position feedback sensors cannot be employed. However, the performance of PEA-driven scanners can be significantly improved without position feedback sensors if an accurate mathematical model with low computational costs is applied to reduce hysteresis and other nonlinear effects. Various methods have been proposed for modeling PEAs, but most of them have limitations in terms of their accuracy and computational efficiencies. In this research, we propose a variant DenseNet-type neural network (NN) model for modeling PEAs in an AFM scanner where position feedback sensors are not available. To improve the performance of this model, the mapping of the forward and backward directions is carried out separately. The experimental results successfully demonstrate the efficacy of the proposed model by reducing the relative root-mean-square (RMS) error to less than 0.1%. Full article
(This article belongs to the Section Actuator Materials)
Show Figures

Figure 1

Figure 1
<p>Experimental setup used to collect data for analyzing the hysteresis of the piezoelectric actuator.</p>
Full article ">Figure 2
<p>Hysteresis curve between the input voltage and the displacement for the X-axis of the homemade scanner.</p>
Full article ">Figure 3
<p>Structure of a fully connected layer model.</p>
Full article ">Figure 4
<p>Mathematical formulation behind an ANN node.</p>
Full article ">Figure 5
<p>Structure of the variant DenseNet-type fully connected model.</p>
Full article ">Figure 6
<p>Identification process.</p>
Full article ">Figure 7
<p>Hysteresis curves fitted with a DenseNet-type neural network for (<b>a</b>) the X-axis and (<b>b</b>) the Y-axis.</p>
Full article ">Figure 8
<p>Comparison of uncompensated/compensated trajectories with desired trajectories (left) and hysteresis error (right) for different driving voltages: (<b>a</b>) 100 V, (<b>b</b>) 75 V, (<b>c</b>) 50 V, (<b>d</b>) 25 V for the X-axis and (<b>e</b>) 100 V, (<b>f</b>) 75 V, (<b>g</b>) 50 V, (<b>h</b>) 25 V for the Y-axis.</p>
Full article ">Figure 9
<p>Tapping mode images of DVD data tracks obtained at 1 Hz with (<b>a</b>) an uncompensated scanner (trace) and (<b>b</b>) an uncompensated scanner (retrace). (<b>c</b>) A line profile for the red and blue lines in (<b>a</b>,<b>b</b>). Tapping mode images of DVD data tracks obtained at 1 Hz with(<b>d</b>) a compensated scanner (trace) and (<b>e</b>) a compensated scanner (retrace). (<b>f</b>) A line profile for the red and blue line in (<b>d</b>,<b>e</b>).</p>
Full article ">Figure 10
<p>Tapping mode images of DVD data tracks obtained at 5 Hz with (<b>a</b>) an uncompensated scanner (trace), (<b>b</b>) an uncompensated scanner (retrace), (<b>c</b>) a compensated scanner (trace), and (<b>d</b>) a compensated scanner (retrace), and at 30 Hz with (<b>e</b>) an uncompensated scanner (trace), (<b>f</b>) an uncompensated scanner (retrace), (<b>g</b>) a compensated scanner (trace), and (<b>h</b>) a compensated scanner (retrace).</p>
Full article ">
10 pages, 4120 KiB  
Article
The Influence of Filler Particle Size on the Strength Properties and Mechanical Energy Dissipation Capacity of Biopoly(Ethylene Terephthalate) BioPET/Eggshell Biocomposites
by Stanisław Kuciel, Karina Rusin-Żurek and Maria Kurańska
Recycling 2024, 9(5), 88; https://doi.org/10.3390/recycling9050088 - 1 Oct 2024
Viewed by 941
Abstract
This work aims to evaluate how the particle size of a waste filler in the form of eggshells changes the mechanical properties of biopoly(ethylene terephthalate) (bioPET). BioPET was modified with three different waste fractions: 1.60–3 mm—large particles; 1.60–1 mm—medium particles; 1 mm–200 μm—small [...] Read more.
This work aims to evaluate how the particle size of a waste filler in the form of eggshells changes the mechanical properties of biopoly(ethylene terephthalate) (bioPET). BioPET was modified with three different waste fractions: 1.60–3 mm—large particles; 1.60–1 mm—medium particles; 1 mm–200 μm—small particles. Waste filler was added to the biopolymer matrix in the amount of 10 wt.%. Static tensile tests, as well as bending and impact tests, were carried out to assess the strength properties of the waste-enriched materials. Dissipation energy changes and relaxation processes were observed and evaluated by means of a low-cycle dynamic test. Waste particles were shown to be an effective modifier of bioPET by increasing its stiffness (all particle sizes) and strength (the smallest ones). Studies of the wetting angle and mechanical energy dissipation in the first hysteresis loops indicate the better adhesion of small particles to the biopolymer and their greater ability to dissipate mechanical energy. Full article
Show Figures

Figure 1

Figure 1
<p>EG particle size distribution curves.</p>
Full article ">Figure 2
<p>Optical microscopy images of modified EG.</p>
Full article ">Figure 3
<p>Tensile strength curves of bioPET composites.</p>
Full article ">Figure 4
<p>SEM images of the biocomposites.</p>
Full article ">Figure 5
<p>Mechanical hysteresis loops for cycle 1 (solid line) and cycle 2 (dashed line).</p>
Full article ">Figure 6
<p>Comparison of mechanical energy dissipation values for cycles 1 and 2.</p>
Full article ">Figure 7
<p>DSC graphs of composite samples: (<b>a</b>) first heating cycle and (<b>b</b>) second heating cycle.</p>
Full article ">
21 pages, 16956 KiB  
Article
Experimental Research on the Seismic Ductility Performance of Wavy Web PEC Beams
by Kejia Yang, Tianyu Lu, Jie Li and Hanzhong Lou
Buildings 2024, 14(10), 3101; https://doi.org/10.3390/buildings14103101 - 27 Sep 2024
Viewed by 487
Abstract
To improve the out-of-plane stability of partially encased composite (PEC) beam webs and enhance the synergy between concrete and section steel, a new type of wavy web PEC beam was designed and fabricated. In this study, the flange thickness and shear–span ratio were [...] Read more.
To improve the out-of-plane stability of partially encased composite (PEC) beam webs and enhance the synergy between concrete and section steel, a new type of wavy web PEC beam was designed and fabricated. In this study, the flange thickness and shear–span ratio were varied as key parameters. Low-cycle reversed loading tests were conducted to investigate the effects of these variables on the load-bearing capacity, failure patterns, deformation capacity, hysteretic energy dissipation capacity, and stiffness degradation of the wavy web PEC beams. Numerical simulations were performed using ABAQUS CAE2023, a finite element analysis (FEA) software, under low-cycle reversed loading conditions. The applicability of the ABAQUS software CAE2023 for the corrugated web PEC beam model was validated by comparing test results with finite element analysis results. A detailed parametric analysis was then carried out using the finite element model to further investigate the mechanical properties of the wavy web PEC beams. The research findings are as follows: the wavy web PEC beams exhibited good ductility; a larger shear–span ratio led to a transition in the failure pattern from shear failure to flexural failure; varying the flange thickness significantly affected the failure location and characteristics; and reducing the flange thickness could limit the propagation of concrete cracks, thereby improving toughness and energy dissipation capacity. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the structure of specimens.</p>
Full article ">Figure 2
<p>Photographs of main steel parts.</p>
Full article ">Figure 3
<p>Specimen loading diagram.</p>
Full article ">Figure 4
<p>Final failure pattern of each component.</p>
Full article ">Figure 5
<p>Load–deflection angle curve of the specimen.</p>
Full article ">Figure 6
<p>Comparison of skeleton curves.</p>
Full article ">Figure 7
<p>Energy equivalence method to determine yield displacement.</p>
Full article ">Figure 8
<p>Schematic diagram of energy dissipation coefficient calculation.</p>
Full article ">Figure 9
<p>Damping coefficient curve.</p>
Full article ">Figure 10
<p>Comparison of stiffness degradation of each component.</p>
Full article ">Figure 11
<p>Comparison of hysteresis curves between simulation and experiment.</p>
Full article ">Figure 12
<p>Comparison of skeleton curves. Comparison of skeleton curves between simulation and experiment.</p>
Full article ">Figure 12 Cont.
<p>Comparison of skeleton curves. Comparison of skeleton curves between simulation and experiment.</p>
Full article ">Figure 13
<p>Stress and damage contours.</p>
Full article ">Figure 13 Cont.
<p>Stress and damage contours.</p>
Full article ">Figure 14
<p>Parametric analysis of peak stress in steel components and the distribution of concrete damage.</p>
Full article ">Figure 14 Cont.
<p>Parametric analysis of peak stress in steel components and the distribution of concrete damage.</p>
Full article ">Figure 15
<p>Influence of flange thickness on load–displacement curves.</p>
Full article ">Figure 16
<p>Influence of shear–span ratio on load–displacement curves.</p>
Full article ">Figure 17
<p>Test component damage index.</p>
Full article ">
22 pages, 5108 KiB  
Article
Enhanced Cyclically Stable Plasticity Model for Multiaxial Behaviour of Magnesium Alloy AZ31 under Low-Cycle Fatigue Conditions
by Aljaž Litrop, Jernej Klemenc, Marko Nagode and Domen Šeruga
Materials 2024, 17(18), 4659; https://doi.org/10.3390/ma17184659 - 23 Sep 2024
Viewed by 785
Abstract
Magnesium alloys, particularly AZ31, are promising materials for the modern automotive industry, offering significant weight savings and environmental benefits. This research focuses on the challenges associated with accurate modelling of multiaxial cyclic plasticity at small strains of AZ31 under low-cycle fatigue conditions. Current [...] Read more.
Magnesium alloys, particularly AZ31, are promising materials for the modern automotive industry, offering significant weight savings and environmental benefits. This research focuses on the challenges associated with accurate modelling of multiaxial cyclic plasticity at small strains of AZ31 under low-cycle fatigue conditions. Current modelling approaches, including crystal plasticity and phenomenological plasticity, have been extensively explored. However, the existing models reach their limits when it comes to capturing the complexity of cyclic plasticity in magnesium alloys, especially under multiaxial loading conditions. To address this gap, a cyclically stable elastoplastic model is proposed that integrates elements from existing models with an enhanced algorithm for updating stresses and hardening parameters, using the hyperbolic tangent function to describe hardening and ensure a stabilised response with closed hysteresis loops for both uniaxial and multiaxial loading. The model is based on a von Mises yield surface and includes a kinematic hardening rule that promises a stable simulation of the response of AZ31 sheets under cyclic loading. Using experimental data from previous studies on AZ31 sheets, the proposed model is optimised and validated. The model shows promising capabilities in simulating the response of AZ31 sheet metal under different loading conditions. It has significant potential to improve the accuracy of fatigue simulations, especially in the context of automotive applications. Full article
(This article belongs to the Section Materials Simulation and Design)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Constitutive model flowchart.</p>
Full article ">Figure 2
<p>Hardening parameters update algorithm.</p>
Full article ">Figure 3
<p>(<b>a</b>) Experimental results of the stabilised hysteresis loop for a uniaxial tensile–compressive load amplitude <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>=</mo> <mo>±</mo> <mn>1</mn> <mo>%</mo> </mrow> </semantics></math>, (<b>b</b>) hardening component <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the compression branch (downwards) of the hysteresis loop, and (<b>c</b>) hardening component <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the tension branch (upwards).</p>
Full article ">Figure 4
<p><math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the compression branch of the hysteresis loop with adapted hyperbolic tangent function.</p>
Full article ">Figure 5
<p><math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> <mfenced separators="|"> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>11</mn> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math> for the tension branch of the hysteresis loop with adapted hyperbolic tangent function.</p>
Full article ">Figure 6
<p>Functions of hardening parameters in dependency of the strain amplitude.</p>
Full article ">Figure 7
<p>Simulated stress–strain responses under uniaxial tensile–compressive loading (dashed lines) compared to the experimental data (dotted lines) for (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±0.5%, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±0.75%, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.0%, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.25%, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math>= ±1.5%, and (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math>= ±1.1%.</p>
Full article ">Figure 8
<p>Simulated stress–strain responses to uniaxial tensile–compressive loading with a strain amplitude <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo>=</mo> <mo>±</mo> <mn>1.0</mn> <mo>%</mo> </mrow> </semantics></math> for different numbers of simulated loading cycles using hardening parameters update algorithm.</p>
Full article ">Figure 9
<p>Simulated stress–strain responses to shear loading (solid lines) with respect to the experimental data (dotted lines) for (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±0.5%, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±0.75%, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.0%, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.25%, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.5%, and (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math>= ±1.1%.</p>
Full article ">Figure 10
<p>Simulated stabilised stress–strain responses under in-phase proportional loading compared to pure uniaxial tension–compression and pure shear loading (solid lines) (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = ±1%, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> = ±0.5%, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = ±1.25%, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> <mo> </mo> </mrow> </msub> </mrow> </semantics></math> = ±0.75%, and (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = ±1.25%, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> <mo> </mo> </mrow> </msub> </mrow> </semantics></math> =±1.25%, and against the experimental data (dotted lines) (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </semantics></math> = ±0.502%, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>γ</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> <mo> </mo> </mrow> </semantics></math> = ±0.654%.</p>
Full article ">
Back to TopTop