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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (75)

Search Parameters:
Keywords = multiple-relaxation-time model

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
34 pages, 4568 KiB  
Review
Nanothermodynamics: There’s Plenty of Room on the Inside
by Ralph V. Chamberlin and Stuart M. Lindsay
Nanomaterials 2024, 14(22), 1828; https://doi.org/10.3390/nano14221828 - 15 Nov 2024
Viewed by 411
Abstract
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical [...] Read more.
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical scaling near ferromagnetic transitions, thermal and dynamic behavior near liquid–glass transitions, and the 1/f-like noise in metal films and qubits. A key feature in several models is to allow separate time steps for distinct conservation laws: one type of step conserves energy and the other conserves momentum (e.g., dipole alignment). This “orthogonal dynamics” explains how the relaxation of a single parameter can exhibit multiple responses such as primary, secondary, and microscopic peaks in the dielectric loss of supercooled liquids, and the crossover in thermal fluctuations from Johnson–Nyquist (white) noise at high frequencies to 1/f-like noise at low frequencies. Nanothermodynamics also provides new insight into three basic questions. First, it gives a novel solution to Gibbs’ paradox for the entropy of the semi-classical ideal gas. Second, it yields the stable equilibrium of Ising’s original model for finite-sized chains of interacting binary degrees of freedom (“spins”). Third, it confronts Loschmidt’s paradox for the arrow of time, showing that an intrinsically irreversible step is required for maximum entropy and the second law of thermodynamics, not only in the thermodynamic limit but also in systems as small as N=2 particles. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Finite-size thermal effects. Inset gives Hill’s fundamental equation of small-system thermodynamics, with a simple (three-energy-level) diagram for each term (adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>]). The first three terms on the right side (black) give the standard ways to increase the total internal energy of a system: add heat (<math display="inline"><semantics> <mrow> <mi>T</mi> <mi>d</mi> <msub> <mrow> <mi>S</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), do work on the system (<math display="inline"><semantics> <mrow> <mo>−</mo> <mi>P</mi> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), or add particles (<math display="inline"><semantics> <mrow> <mi>μ</mi> <mi>d</mi> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>). The fourth term (red) contains finite-size effects (surface states, length-scale terms, fluctuations, etc.) that change the width of the levels when the number of subdivisions changes if the subdivision potential is nonzero (<math display="inline"><semantics> <mrow> <mo>ℇ</mo> <mo>≠</mo> <mn>0</mn> </mrow> </semantics></math>). The main figure shows how free energy might change with the number of subdivisions, from <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in the thermodynamic limit of no subdivisions (<math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) to <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">n</mi> </mrow> </semantics></math> in the nanothermodynamic limit for stable equilibrium of subsystems inside bulk samples (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">ℇ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Schematic representation of various multiplicities. A canonical system (<b>top</b>) has two indistinguishable particles that may be on the left side (L), right side (R), or opposite sides. There is only one way to subdivide this system into canonical subsystems (<b>middle</b>), but there are many ways to subdivide it into nanocanonical subsystems (<b>bottom</b>). Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>Sketch showing two solutions to Gibbs’ paradox for combining two types of particles: X’s (blue) and O’s (red). (<b>A</b>–<b>C</b>) Canonical ensemble, where all particles of the same type are indistinguishable over all distances. (<b>D</b>–<b>F</b>) Nanocanonical ensemble, comprised of nanoscale subsystems, where similar particles can be distinguished by their location when in different subsystems. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 4
<p>Sketch showing a stable solution of the 1D Ising model at a given <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math>. Ten spins are in the chain. Each spin may be up or down. Each interaction between neighboring spins may be low energy (<math display="inline"><semantics> <mrow> <mo>●</mo> </mrow> </semantics></math>), high energy (<b>X</b>), or a no-energy “break” (<b>O</b>) in the interaction.</p>
Full article ">Figure 5
<p>(<b>D</b>) Temperature dependence of the effective scaling exponent from data (symbols) and models (lines) sketched in (<b>A</b>–<b>C</b>). Each red box encloses a separate set of spins that can be treated using mean-field theory. (<b>A</b>) Standard mean-field theory yields <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (dotted line in (<b>D</b>)). (<b>B</b>) Simulations of the standard Ising model yield a monotonic increase in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> with decreasing <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> (dashed line in (<b>D</b>)). (<b>C</b>) The mean-field cluster model yields non-monotonic behavior in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> (solid lines in (<b>D</b>)), similar to measurements on EuO (circles) and Gd (squares). Difficulty in determining <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math> yields uncertainty as <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>→</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>, but not for <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> <mo>[</mo> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>T</mi> <mo>−</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>)</mo> <mo>/</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>]</mo> </mrow> </mrow> <mo>&gt;</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> where <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> of the standard Ising model shows only gradual and monotonic behavior, unlike the measurements. Adapted from [<a href="#B26-nanomaterials-14-01828" class="html-bibr">26</a>].</p>
Full article ">Figure 6
<p>Log-log plot of frequency-dependent loss from the orthogonal Ising model. The loss is deduced from the power spectral density (PSD) using the fluctuation-dissipation theorem. The frequency is normalized by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> to put the microscopic peak at <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>f</mi> <mo>/</mo> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> <mo>~</mo> <mn>12</mn> </mrow> </semantics></math>. Simulations are made on subsystems of two sizes, each at two temperatures, as given in the legends. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 7
<p>Primary response time of glycerol. Abscissa is inverse temperature, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>/</mo> <mi>T</mi> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> is the mean-field critical temperature. The ordinate in (<b>A</b>) is <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, and in (<b>B</b>) it comes from a type of Stickel plot [<a href="#B84-nanomaterials-14-01828" class="html-bibr">84</a>] utilizing finite differences of <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, which removes the prefactor and linearize the VFT2 function. Symbols are from measurements (Stickel [<a href="#B85-nanomaterials-14-01828" class="html-bibr">85</a>]). Various lines are from the VFT2 function Equation (6) (black), VFT function (red), and MYEGA function (blue) [<a href="#B86-nanomaterials-14-01828" class="html-bibr">86</a>]. The inset is a sketch of a simple free-energy diagram, containing two minima separated by a barrier. Primary response in the orthogonal Ising model involves fluctuations in energy that open pathways between the minima. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 8
<p>1/<span class="html-italic">f</span>-like noise from maintaining maximum entropy during equilibrium fluctuations. (<b>A</b>–<b>E</b>) Sketch of all distinct configurations of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> spins, arranged in order of decreasing alignment from <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> (<b>top</b>) to <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> (<b>bottom</b>). The multiplicity for the alignment entropy of the subsystem comes from the number of configurations in each box. (<b>F</b>) Temperature-dependent exponent for noise that varies as a function of frequency, <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>S</mi> <mi>D</mi> <mo>(</mo> <mi>f</mi> <mo>)</mo> <mo>∝</mo> <mn>1</mn> <mo>/</mo> <msup> <mrow> <mi>f</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msup> </mrow> </semantics></math>, with the abscissa normalized by <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>. Solid symbols (color) are from measurements [<a href="#B88-nanomaterials-14-01828" class="html-bibr">88</a>] of noise in thin films for the metals given in the legend. Open symbols (black) are from simulations of a 3D Ising subsystems having <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>27</mn> </mrow> </semantics></math> spins with dynamics utilizing a local bath to maintain maximum entropy during fluctuations in alignment. Solid line is the best linear fit to the simulations, weighted by the inverse variance of each point. Dashed line is from a random fluctuation model [<a href="#B89-nanomaterials-14-01828" class="html-bibr">89</a>]. Adapted from [<a href="#B60-nanomaterials-14-01828" class="html-bibr">60</a>].</p>
Full article ">Figure 9
<p>Influence of energy on the amplitude of alignment fluctuations via orthogonal dynamics. (<b>A</b>–<b>E</b>) Configurations of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> interactions, arranged in order of decreasing energy. (<b>F</b>) Simulation of energy (<math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> </mrow> </semantics></math>, red) and magnetization (<math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math>, black) as a function of time for the 1D Ising model containing <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> interactions, with a local bath to maintain maximum entropy. Note how the amplitude of fluctuations in <math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math> tends to be slightly larger when <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> <mo>&lt;</mo> <mo>−</mo> <mn>0.3</mn> </mrow> </semantics></math>. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 10
<p>Noise power spectral densities from simulations (lines) and measurements (symbols). Solid lines are from fluctuations in alignment of 1D chains of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> Ising spins using orthogonal dynamics while maintaining maximum entropy. Note that <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> (blue) is a small enough subsystem to show separate Lorentzians in a 1/<span class="html-italic">f</span>-like spectrum, while <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> (red) is large enough to show a crossover from white noise at high frequencies (dotted) to 1/<span class="html-italic">f</span>-like noise at low frequencies with an exponent of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.92</mn> </mrow> </semantics></math> (dashed). Symbols are from measurements of flux noise (solid) and tunnel-coupling noise (open) in a qubit [<a href="#B92-nanomaterials-14-01828" class="html-bibr">92</a>]. Each set of measurements has been shifted in amplitude and frequency to match the simulations. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 11
<p>Time dependence of entropies per particle (<b>A</b>–<b>E</b>) and inverse effective temperatures (<b>F</b>). Simulations utilize a Creutz-like model of 1D Ising-like spins coupled to a <span class="html-italic">ke</span> bath of Einstein oscillators. Top three left-side graphs show the time-dependence of <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>/</mo> <mo>(</mo> <mi>N</mi> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> for the spins (<b>C</b>), <span class="html-italic">ke</span> bath (<b>B</b>), and their sum (<b>A</b>) in a large system, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>. Symbols come from first averaging 10,000 sweeps, then averaging three separate simulations of each type, with error bars visible if larger than the symbol size. A simulation with irreversible dynamics (red circles) precedes every simulation with reversible dynamics (black squares). Thus, the total entropy always decreases when the dynamics becomes reversible, as indicated by the orange arrow in (<b>A</b>). Furthermore, when the rate of break-change attempts is reduced to 1/10 the rate of spin-change attempts (middle third of every simulation), reversible simulations have an entropy that depends on the dynamics. Right-side graphs show the total entropies, as in (<b>A</b>) but without time-averaging, over a greatly expanded time scale. Here the differences between reversible (black) and irreversible (red) behavior are clearly visible at the start (<b>D</b>) and end (<b>E</b>) of the simulations. The inset shows corresponding differences in the power-spectral densities of the simulations. Symbols in (<b>F</b>) give the logarithm of the ratio of probabilities of neighboring energy levels in the <span class="html-italic">ke</span> bath, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> <mo>/</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> <mo>+</mo> <mn>1</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (squares), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (circles), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> (up triangles), and <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (down triangles). These values are proportional to the difference in inverse effective temperature of the adjacent levels. A single temperature applies only to irreversible dynamics in the thermodynamic limit (red), not for reversible dynamics in this limit (black) nor for irreversible dynamics of small subsystems, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>128</mn> </mrow> </semantics></math> (green). Adapted from [<a href="#B16-nanomaterials-14-01828" class="html-bibr">16</a>].</p>
Full article ">Figure 12
<p>Fluctuations in potential energy from MD simulations of Lennard–Jones crystals. Main figure shows normalized <span class="html-italic">pe</span> fluctuations for blocks of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>32</mn> </mrow> </semantics></math> atoms in a system of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>442,368</mn> </mrow> </semantics></math> atoms as a function interaction cutoff radius, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>, at three temperatures given in the legend. Note that the data (open symbols) tend to be relatively constant (independent of <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>) when interactions are robustly harmonic, having interaction between nearest-neighbor atoms only, <math display="inline"><semantics> <mrow> <mn>1.12</mn> <mo>≈</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≤</mo> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>≤</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>4</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≈</mo> <mn>1.59</mn> </mrow> </semantics></math>. Insets show the time dependence of energy autocorrelations in blocks (black squares) and energy correlations between nearest-neighbor blocks (red circles). Simulations are made at <math display="inline"><semantics> <mrow> <mi>k</mi> <mi>T</mi> <mo>/</mo> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mn>0.0005</mn> </mrow> </semantics></math> for blocks containing a single unit cell of the crystal, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>. The lower inset shows that neighboring blocks are positively correlated when all atoms have robustly harmonic interactions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>), while the upper inset (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>2.0</mn> </mrow> </semantics></math>) shows that neighboring blocks are anticorrelated when interactions include second-neighbor atoms that are anharmonic. Adapted from [<a href="#B15-nanomaterials-14-01828" class="html-bibr">15</a>] with permission from Elsevier.</p>
Full article ">
22 pages, 4076 KiB  
Article
Simulation and Analysis of the Loading, Relaxation, and Recovery Behavior of Polyethylene and Its Pipes
by Furui Shi and P.-Y. Ben Jar
Polymers 2024, 16(22), 3153; https://doi.org/10.3390/polym16223153 - 12 Nov 2024
Viewed by 693
Abstract
Spring–dashpot models have long been used to simulate the mechanical behavior of polymers, but their usefulness is limited because multiple model parameter values can reproduce the experimental data. In view of this limitation, this study explores the possibility of improving uniqueness of parameter [...] Read more.
Spring–dashpot models have long been used to simulate the mechanical behavior of polymers, but their usefulness is limited because multiple model parameter values can reproduce the experimental data. In view of this limitation, this study explores the possibility of improving uniqueness of parameter values so that the parameters can be used to establish the relationship between deformation and microstructural changes. An approach was developed based on stress during the loading, relaxation, and recovery of polyethylene. In total, 1000 sets of parameter values were determined for fitting the data from the relaxation stages with a discrepancy within 0.08 MPa. Despite a small discrepancy, the 1000 sets showed a wide range of variation, but one model parameter, σv,L0, followed two distinct paths rather than random distribution. The five selected sets of parameter values with discrepancies below 0.04 MPa were found to be highly consistent, except for the characteristic relaxation time. Therefore, this study concludes that the uniqueness of model parameter values can be improved to characterize the mechanical behavior of polyethylene. This approach then determined the quasi-static stress of four polyethylene pipes, which showed that these pipes had very close quasi-static stress. This indicates that the uniqueness of the parameter values can be improved for the spring–dashpot model, enabling further study using spring–dashpot models to characterize polyethylene’s microstructural changes during deformation. Full article
(This article belongs to the Special Issue Polymers Physics: From Theory to Experimental Applications)
Show Figures

Figure 1

Figure 1
<p>Specimens used in the RR tests: (<b>a</b>) dimensions of cylindrical specimen, (<b>b</b>) cylindrical specimen, (<b>c</b>) dimensions of NPR specimen (PE-Xa pipe as an example), and (<b>d</b>) NPR specimen. All units are in millimeters.</p>
Full article ">Figure 2
<p>Three-branch spring–dashpot model used in this study.</p>
Full article ">Figure 3
<p>Procedure for the determination of fitting parameters in the relaxation, recovery, and loading stages of RR tests.</p>
Full article ">Figure 4
<p>The 1000 sets of parameter values for simulation at the relaxation stages of different deformation levels in one RR test of HDPE-b: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>S</mi> </mrow> </msub> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>, and (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>q</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>. Different colors at one stroke are used to indicate the 1000 sets of parameter values.</p>
Full article ">Figure 5
<p>A two-path pattern of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> as a function of stroke for NPR specimens based on 1000 sets of parameter values: (<b>a</b>) PE-Xa, (<b>b</b>) PE2708, (<b>c</b>) PE4710-yellow, and (<b>d</b>) PE4710-black pipes. Different colors at one stroke are used to indicate the 1000 sets of parameter values.</p>
Full article ">Figure 6
<p>Best five sets of parameter values (in open red circles) selected from 1000 sets for the simulation of stress variation at the relaxation stages of HDPE-b and the corresponding <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>q</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>: (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math>, (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>S</mi> </mrow> </msub> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mn>0</mn> <mo>,</mo> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>e</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math>, (<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>, and (<b>g</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>q</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>. Different colors at one stroke are used to indicate the five best sets of parameter values.</p>
Full article ">Figure 7
<p>Simulation of stress change at relaxation stages of different strokes for HDPE-b using the fitting parameter values in <a href="#polymers-16-03153-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 8
<p><math display="inline"><semantics> <mrow> <msub> <mrow> <mi>K</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>K</mi> </mrow> <mrow> <mi>v</mi> <mo>,</mo> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math> as a function of stroke of HDPE-b.</p>
Full article ">Figure 9
<p>Summary RR test results for NPR specimens: (<b>a</b>) applied stress at the onset of relaxation,<math display="inline"><semantics> <mrow> <mo> </mo> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>A</mi> </mrow> </msub> </mrow> </semantics></math>(0), and (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>q</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">
32 pages, 940 KiB  
Article
Modeling and Optimization of the Inland Container Transportation Problem Considering Multi-Size Containers, Fuel Consumption, and Carbon Emissions
by Yujian Song and Yuting Zhang
Processes 2024, 12(10), 2231; https://doi.org/10.3390/pr12102231 - 13 Oct 2024
Viewed by 926
Abstract
This paper investigates the inland container transportation problem with a focus on multi-size containers, fuel consumption, and carbon emissions. To reflect a more realistic situation, the depot’s initial inventory of empty containers is also taken into consideration. To linearly model the constraints imposed [...] Read more.
This paper investigates the inland container transportation problem with a focus on multi-size containers, fuel consumption, and carbon emissions. To reflect a more realistic situation, the depot’s initial inventory of empty containers is also taken into consideration. To linearly model the constraints imposed by the multiple container sizes and the limited number of empty containers, a novel graphical representation is presented for the problem. Based on the graphical representation, a mixed-integer programming model is presented to minimize the total transportation cost, which includes fixed, fuel, and carbon emission costs. To efficiently solve the model, a tailored branch-and-price algorithm is designed, which is enhanced by improvement schemes including a heuristic label-setting algorithm, decremental state-space relaxation, and the introduction of a high-quality upper bound. Results from a series of computational experiments on randomly generated instances demonstrate that (1) the proposed branch-and-price algorithm demonstrates a superior performance compared to the tabu search algorithm and the genetic algorithm; (2) each additional empty container in the depot reduces the total transportation cost by less than 1%, with a diminishing marginal effect; (3) the rational configuration of different types of trucks improves scheduling flexibility and reduces fuel and carbon emission costs as well as the overall transportation cost; and (4) extending customer time windows also contributes to lower the total transportation cost. These findings not only deepen the theoretical understanding of inland container transportation optimization but also provide valuable insights for logistics companies and policymakers to improve efficiency and implement more sustainable operational practices. Additionally, our research paves the way for future investigations into the integration of dynamic factors and emerging technologies in this field. Full article
(This article belongs to the Section Sustainable Processes)
Show Figures

Figure 1

Figure 1
<p>The flowchart of this paper.</p>
Full article ">Figure 2
<p>Visualization of transportation operations for 20 ft containers.</p>
Full article ">Figure 3
<p>An illustrative example.</p>
Full article ">Figure 4
<p>Node definitions for container transportation requests.</p>
Full article ">Figure 5
<p>Nodes representing empty container storage and retrieval from the depot.</p>
Full article ">Figure 6
<p>Workflow diagram for the proposed branch-and-price algorithm.</p>
Full article ">Figure 7
<p>Time window branching.</p>
Full article ">Figure 8
<p>Comparison of Gap values for different algorithms.</p>
Full article ">Figure 9
<p>Relationship between total cost and the initial number of 40 ft empty containers at the depot.</p>
Full article ">Figure 10
<p>Heatmap of total transportation cost under different truck fleet configurations.</p>
Full article ">Figure 11
<p>Heatmap of fuel and carbon emission costs under different truck fleet configurations.</p>
Full article ">Figure 12
<p>Relationship between time window length and costs.</p>
Full article ">
17 pages, 4525 KiB  
Article
An Investigation of Thermomechanical Behavior in Laser Hot Wire Directed Energy Deposition of NAB: Finite Element Analysis and Experimental Validation
by Glenn W. Hatala, Edward Reutzel and Qian Wang
Metals 2024, 14(10), 1143; https://doi.org/10.3390/met14101143 - 8 Oct 2024
Viewed by 575
Abstract
Laser Hot Wire (LHW) Directed Energy Deposition (DED) Additive Manufacturing (AM) processes are capable of manufacturing parts with a high deposition rate. There is a growing research interest in replacing large cast Nickel Aluminum Bronze (NAB) components using LHW DED processes for maritime [...] Read more.
Laser Hot Wire (LHW) Directed Energy Deposition (DED) Additive Manufacturing (AM) processes are capable of manufacturing parts with a high deposition rate. There is a growing research interest in replacing large cast Nickel Aluminum Bronze (NAB) components using LHW DED processes for maritime applications. Understanding thermomechanical behavior during LHW DED of NAB is a critical step towards the production of high-quality NAB parts with desired performance and properties. In this paper, finite element simulations are first used to predict the thermomechanical time histories during LHW DED of NAB test coupons with an increasing geometric complexity, including single-layer and multilayer depositions. Simulation results are experimentally validated through in situ measurements of temperatures at multiple locations in the substrate as well as displacement at the free end of the substrate during and immediately following the deposition process. The results in this paper demonstrate that the finite element predictions have good agreement with the experimental measurements of both temperature and distortion history. The maximum prediction error for temperature is 5% for single-layer samples and 6% for multilayer samples, while the distortion prediction error is about 12% for single-layer samples and less than 4% for multilayer samples. In addition, this study shows the effectiveness of including a stress relaxation temperature at 500 °C during FE modeling to allow for better prediction of the low cross-layer accumulation of distortion in multilayer deposition of NAB. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup, including robotic LHW DED system and clamped substrate.</p>
Full article ">Figure 2
<p>Single-layer part dimensions. All measurements are in mm; T1–T3 denote the thermocouples at three locations on the substrate.</p>
Full article ">Figure 3
<p>In situ sensing setup for thermal and mechanical measurements.</p>
Full article ">Figure 4
<p>Dilatometry curve illustrating the volumetric change in microstructural transformations for Cu-12Al during a heating and cooling cycle [<a href="#B28-metals-14-01143" class="html-bibr">28</a>].</p>
Full article ">Figure 5
<p>Meshing used in the simulation of a multilayer build.</p>
Full article ">Figure 6
<p>Images of sample depositions.</p>
Full article ">Figure 7
<p>Single-layer prediction versus experimental measurements, where the coefficient of laser absorptivity <math display="inline"><semantics> <mi>η</mi> </semantics></math> = 0.19 and the convection coefficient = 35 W/(m<sup>2</sup> K): (<b>a</b>) temperature history at T1–T3 and (<b>b</b>) displacement history at LDS. The gray shaded areas indicate the laser-active time in depositing each bead.</p>
Full article ">Figure 8
<p>Multilayer prediction versus experimental measurements, where the coefficient of laser absorptivity <math display="inline"><semantics> <mi>η</mi> </semantics></math> = 0.19 and the convection coefficient = 35 W/(m<sup>2</sup> K): (<b>a</b>) temperature history at T1–T3 and (<b>b</b>) displacement history at LDS. The gray shaded areas indicate the laser-active time in depositing each bead.</p>
Full article ">
18 pages, 5215 KiB  
Article
Cascaded Vehicle State Estimation Method of 4WIDEVs Considering System Delay and Noise
by Zibin Yang, Xiang Liu and Qiu Xia
World Electr. Veh. J. 2024, 15(10), 454; https://doi.org/10.3390/wevj15100454 - 7 Oct 2024
Viewed by 684
Abstract
Considering the negative effects of time delay and noise on vehicle state estimation, a cascaded estimation means for the vehicle sideslip angle is proposed utilizing the ODUKF algorithm. To achieve strong-correlation decoupling between state variables and model interference of the EDWM, an augmented [...] Read more.
Considering the negative effects of time delay and noise on vehicle state estimation, a cascaded estimation means for the vehicle sideslip angle is proposed utilizing the ODUKF algorithm. To achieve strong-correlation decoupling between state variables and model interference of the EDWM, an augmented EDWM was constructed by introducing the tire relaxation length dynamic equation, which enables the precise model relationship between the longitudinal and transverse tire force relaxation length to be constructed while also achieving the decoupling of the system state from the unknown input. To achieve a vehicle driving state estimation, a hierarchical estimation architecture was adopted to design a cascading estimation method for the vehicle driving state. By using tire force estimation values as input for the vehicle driving state estimation, the required vehicle body postures can be estimated. At the same time, facing the problems of system delay and noise, an estimator derived from the ODUKF is designed by combining the model and cascade estimation strategy. The simulation comparative analysis and quantitative statistical results under multiple operating conditions provide evidence that the developed means effectively heighten the estimation accurateness and real-time performance while considering system time delay and noise. Full article
Show Figures

Figure 1

Figure 1
<p>Cascaded estimation strategy.</p>
Full article ">Figure 2
<p>Steering wheel angle in case 1.</p>
Full article ">Figure 3
<p>Results of <span class="html-italic">F<sub>x</sub></span> in case 1.</p>
Full article ">Figure 4
<p>Results of <span class="html-italic">F<sub>y</sub></span> in case 1.</p>
Full article ">Figure 5
<p>Vehicle body postures in case 1. (<b>a</b>) <span class="html-italic">v</span>, (<b>b</b>) <span class="html-italic">u</span>, and (<b>c</b>) <span class="html-italic">β</span>.</p>
Full article ">Figure 6
<p>States in case 2. (<b>a</b>) Steering wheel angle and (<b>b</b>) vehicle speed.</p>
Full article ">Figure 7
<p>Results of <span class="html-italic">F<sub>x</sub></span> in case 2.</p>
Full article ">Figure 8
<p>Results of <span class="html-italic">F<sub>y</sub></span> in case 2.</p>
Full article ">Figure 9
<p>Vehicle body postures in case 2. (<b>a</b>) <span class="html-italic">v</span>, (<b>b</b>) <span class="html-italic">u</span>, and (<b>c</b>) <span class="html-italic">β</span>.</p>
Full article ">
20 pages, 17123 KiB  
Article
A Physics-Informed Neural Network Based on the Boltzmann Equation with Multiple-Relaxation-Time Collision Operators
by Zhixiang Liu, Chenkai Zhang, Wenhao Zhu and Dongmei Huang
Axioms 2024, 13(9), 588; https://doi.org/10.3390/axioms13090588 - 29 Aug 2024
Viewed by 748
Abstract
The Boltzmann equation with multiple-relaxation-time (MRT) collision operators has been widely employed in kinetic theory to describe the behavior of gases and liquids at the macro-level. Given the successful development of deep learning and the availability of data analytic tools, it is a [...] Read more.
The Boltzmann equation with multiple-relaxation-time (MRT) collision operators has been widely employed in kinetic theory to describe the behavior of gases and liquids at the macro-level. Given the successful development of deep learning and the availability of data analytic tools, it is a feasible idea to try to solve the Boltzmann-MRT equation using a neural network-based method. Based on the canonical polyadic decomposition, a new physics-informed neural network describing the Boltzmann-MRT equation, named the network for MRT collision (NMRT), is proposed in this paper for solving the Boltzmann-MRT equation. The method of tensor decomposition in the Boltzmann-MRT equation is utilized to combine the collision matrix with discrete distribution functions within the moment space. Multiscale modeling is adopted to accelerate the convergence of high frequencies for the equations. The micro–macro decomposition method is applied to improve learning efficiency. The problem-dependent loss function is proposed to balance the weight of the function for different conditions at different velocities. These strategies will greatly improve the accuracy of the network. The numerical experiments are tested, including the advection–diffusion problem and the wave propagation problem. The results of the numerical simulation show that the network-based method can obtain a measure of accuracy at O103. Full article
(This article belongs to the Section Mathematical Analysis)
Show Figures

Figure 1

Figure 1
<p>Network architecture. <math display="inline"><semantics> <mi mathvariant="bold">x</mi> </semantics></math> is spatial space and t is time, which all are the inputs of the network. The Monte Carlo method is used to create a dataset. Multiscale modeling and canonical polyadic decomposition are adopted in the neural network.</p>
Full article ">Figure 2
<p>Compositions of the loss function. <math display="inline"><semantics> <mi mathvariant="bold">x</mi> </semantics></math> is spatial space and t is time, which are inputs to the network. The Monte Carlo method is used to create a dataset. The loss function is made up of three parts: IC loss, BC loss, and PDE loss. The problem-dependent weights for the three loss parts are adopted.</p>
Full article ">Figure 3
<p>Numerical solution of advection–diffusion problem using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mn>0.1</mn> <mo>,</mo> <mn>1.0</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math>. The numerical solution of NMRT is the solid line, and the reference solution is the dashed line. (<b>a</b>) s = 0.01, t = 0; (<b>b</b>) s = 0.01, t = 0.1; (<b>c</b>) s = 0.1, t = 0; (<b>d</b>) s = 0.1, t = 0.1; (<b>e</b>) s = 1.0, t = 0; (<b>f</b>) s = 1.0, t = 0.1.</p>
Full article ">Figure 3 Cont.
<p>Numerical solution of advection–diffusion problem using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mn>0.1</mn> <mo>,</mo> <mn>1.0</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math>. The numerical solution of NMRT is the solid line, and the reference solution is the dashed line. (<b>a</b>) s = 0.01, t = 0; (<b>b</b>) s = 0.01, t = 0.1; (<b>c</b>) s = 0.1, t = 0; (<b>d</b>) s = 0.1, t = 0.1; (<b>e</b>) s = 1.0, t = 0; (<b>f</b>) s = 1.0, t = 0.1.</p>
Full article ">Figure 4
<p>Numerical solution of wave propagation problem using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>0.01</mn> <mo>,</mo> <mn>0.1</mn> <mo>,</mo> <mn>1.0</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.0</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math>. The numerical solution of NMRT is the solid line, and the reference solution is the dashed line. (<b>a</b>) s = 0.01, t = 0; (<b>b</b>) s = 0.01, t = 0.1; (<b>c</b>) s = 0.1, t = 0; (<b>d</b>) s = 0.1, t = 0.1; (<b>e</b>) s = 1.0, t = 0; (<b>f</b>) s = 1.0, t = 0.1.</p>
Full article ">Figure 5
<p>Numerical solution of the wave propagation problem in two-dimensional scenarios using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>. The first row corresponds to the density <math display="inline"><semantics> <mi>ρ</mi> </semantics></math>, the second column corresponds to the macroscopic velocity <span class="html-italic">u</span>, and the last column corresponds to the temperature <span class="html-italic">T</span>. The numerical solution of NMRT is the solid line, and the result of FSM is the dashed line. (<b>a</b>) <span class="html-italic">s</span> = 0.01, <span class="html-italic">t</span> = 0.1; (<b>b</b>) <span class="html-italic">s</span> = 0.01, <span class="html-italic">t</span> = 0.1; (<b>c</b>) <span class="html-italic">s</span> = 0.01, <span class="html-italic">t</span> = 0.1.</p>
Full article ">Figure 6
<p>Numerical solution of wave propagation problem in two-dimensional scenarios using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>. The first row corresponds to the density <math display="inline"><semantics> <mi>ρ</mi> </semantics></math>, the second column corresponds to the macroscopic velocity <span class="html-italic">u</span>, and the last column corresponds to the temperature <span class="html-italic">T</span>. The numerical solution of NMRT is the solid line, and the result of FSM is the dashed line. (<b>a</b>) <span class="html-italic">s</span> = 0.1, <span class="html-italic">t</span> = 0.1; (<b>b</b>) <span class="html-italic">s</span> = 0.1, <span class="html-italic">t</span> = 0.1; (<b>c</b>) <span class="html-italic">s</span> = 0.1, <span class="html-italic">t</span> = 0.1.</p>
Full article ">Figure 7
<p>Numerical solution of the wave propagation problem in two-dimensional scenarios using NMRT for <math display="inline"><semantics> <mrow> <mi>s</mi> <mo>=</mo> <mn>1.0</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>0.1</mn> </mrow> </semantics></math>. The first row corresponds to the density <math display="inline"><semantics> <mi>ρ</mi> </semantics></math>, the second column corresponds to the macroscopic velocity <span class="html-italic">u</span>, and the last column corresponds to the temperature <span class="html-italic">T</span>. The numerical solution of NMRT is the solid line, and the result of FSM is the dashed line. (<b>a</b>) <span class="html-italic">s</span> = 1.0, <span class="html-italic">t</span> = 0.1; (<b>b</b>) <span class="html-italic">s</span> = 1.0, <span class="html-italic">t</span> = 0.1; (<b>c</b>) <span class="html-italic">s</span> = 1.0, <span class="html-italic">t</span> = 0.1.</p>
Full article ">
14 pages, 292 KiB  
Article
A Non-Linear Optimization Model for the Multi-Depot Multi-Supplier Vehicle Routing Problem with Relaxed Time Windows
by Herman Mawengkang, Muhammad Romi Syahputra, Sutarman Sutarman and Abdellah Salhi
Vehicles 2024, 6(3), 1482-1495; https://doi.org/10.3390/vehicles6030070 - 29 Aug 2024
Viewed by 1007
Abstract
In the realm of supply chain logistics, the Multi-Depot Multi-Supplier Vehicle Routing Problem (MDMSVRP) poses a significant challenge in optimizing the transportation process to minimize costs and enhance operational efficiency. This problem involves determining the most cost-effective routes for a fleet of vehicles [...] Read more.
In the realm of supply chain logistics, the Multi-Depot Multi-Supplier Vehicle Routing Problem (MDMSVRP) poses a significant challenge in optimizing the transportation process to minimize costs and enhance operational efficiency. This problem involves determining the most cost-effective routes for a fleet of vehicles to deliver goods from multiple suppliers to multiple depots, considering various constraints and non-linear relationships. The routing problem (RP) is a critical element of many logistics systems that involve the routing and scheduling of vehicles from a depot to a set of customer nodes. One of the most studied versions of the RP is the Vehicle Routing Problem with Time Windows (VRPTW), in which each customer must be visited at certain time intervals, called time windows. In this paper, it is considered that there are multiple depots (supply centers) and multiple suppliers, along with a fleet of vehicles. The goal is to efficiently plan routes for these vehicles to deliver goods from the suppliers to various customers while considering relaxed time windows. This research is intended to establish a new relaxation scheme that relaxes the time window constraints in order to lead to feasible and good solutions. In addition, this study develops a discrete optimization model as an alternative model for the time-dependent VRPTW involving multi-suppliers. This research also develops a metaheuristic algorithm with an initial solution that is determined through time window relaxation. Full article
16 pages, 1702 KiB  
Article
Influence of Local Thermodynamic Non-Equilibrium to Photothermally Induced Acoustic Response of Complex Systems
by Slobodanka Galovic, Aleksa I. Djordjevic, Bojan Z. Kovacevic, Katarina Lj. Djordjevic and Dalibor Chevizovich
Fractal Fract. 2024, 8(7), 399; https://doi.org/10.3390/fractalfract8070399 - 3 Jul 2024
Viewed by 871
Abstract
In this paper, the time-resolved model of the photoacoustic signal for samples with a complex inner structure is derived including local non-equilibrium of structural elements with multiple degrees of freedom, i.e., structural entropy of the system. The local non-equilibrium is taken into account [...] Read more.
In this paper, the time-resolved model of the photoacoustic signal for samples with a complex inner structure is derived including local non-equilibrium of structural elements with multiple degrees of freedom, i.e., structural entropy of the system. The local non-equilibrium is taken into account through the fractional operator. By analyzing the model for two types of time-dependent excitation, a very short pulse and a very long pulse, it is shown that the rates of non-equilibrium relaxations in complex samples can be measured by applying the derived model and time-domain measurements. Limitations of the model and further directions of its development are discussed. Full article
Show Figures

Figure 1

Figure 1
<p>Geometry of the problem.</p>
Full article ">Figure 2
<p>Photoacoustic signal induced by very short optical pulse (Equations (48) and (50)) for (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>&gt;</mo> <mn>0.3</mn> </mrow> </semantics></math> (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>&lt;</mo> <mn>0.3</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>The PA signals excited by very long optical pulse (described by Equations (49) and (51)) for <math display="inline"><semantics> <mrow> <mi>a</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and several values of the fractional exponent <math display="inline"><semantics> <mi>υ</mi> </semantics></math> from the range <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <mn>0</mn> <mo>,</mo> <mn>1</mn> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>The dependence of PA signal on sample thickness (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math>, short pulse excitation; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, short pulse excitation, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>=</mo> <mn>0.5</mn> </mrow> </semantics></math>, long pulse excitation; and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>υ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>, long pulse excitation.</p>
Full article ">
17 pages, 5725 KiB  
Article
A GPU-Implemented Lattice Boltzmann Model for Large Eddy Simulation of Turbulent Flows in and around Forest Shelterbelts
by Yansen Wang, Xiping Zeng, Jonathan Decker and Leelinda Dawson
Atmosphere 2024, 15(6), 735; https://doi.org/10.3390/atmos15060735 - 20 Jun 2024
Cited by 1 | Viewed by 852
Abstract
Using porous wind barriers for the microclimate modification of agricultural lands, urban areas, and surrounding roads is a ubiquitous practice. This study establishes a new method for numerically modeling the turbulent flow in and around forest shelterbelts using an advanced multiple-relaxation-time lattice Boltzmann [...] Read more.
Using porous wind barriers for the microclimate modification of agricultural lands, urban areas, and surrounding roads is a ubiquitous practice. This study establishes a new method for numerically modeling the turbulent flow in and around forest shelterbelts using an advanced multiple-relaxation-time lattice Boltzmann model (MRTLBM). A detailed description is presented for a large eddy simulation (LES) of turbulent winds by implementing barrier element drag force in the MRTLBM framework. The model results for a forest shelterbelt are compared with a field observational dataset. The study indicated that our implementation of drag force in MRTLBM is an accurate method for modeling turbulent flows in and around forest patches. Sensitivity analyses of turbulent flow related to the shelterbelt structure parameters and wind directions are also carried out. The analysis indicated that the optimal wind shelter effect in reducing the mean wind speed and turbulent kinetic energy is maximized using a narrow, medium porosity shelterbelt, with the wind direction perpendicular to the shelterbelt. These conclusions are in agreement with other observational and modeling studies. Finally, the computational time of a central processing unit (CPU) and graphics processing unit (GPU) was compared for a large domain with 25 million grids to demonstrate the MRTLBM advantage of LES in regards to computational speed with a mixed forest and building environment. The GPU is approximately 300 times faster than a CPU, and real-time simulation for this large domain is achieved using the Nvidia V100 GPU. Full article
(This article belongs to the Section Atmospheric Techniques, Instruments, and Modeling)
Show Figures

Figure 1

Figure 1
<p>The observational setup in Kurotani et al. [<a href="#B6-atmosphere-15-00735" class="html-bibr">6</a>]. The wind was perpendicular to the shelterbelt, denoted in green. P<sub>0</sub> is the reference inflow profile 35 m upwind from the shelterbelt. P<sub>1</sub>–P<sub>5</sub> profiles were on the lee side of the shelterbelt. Blue dots are the 3D sonic observational points at the denoted Z heights from the ground.</p>
Full article ">Figure 2
<p>The simulation domain setup. The observational profiles are located at the symmetric xy plane, as denoted by P<sub>0</sub>–P<sub>5</sub>.</p>
Full article ">Figure 3
<p>The inflow mean wind profile (<b>left panel</b>), U, and the TKE (<b>right panel</b>) at the location of reference tower (P<sub>0</sub>). The circles indicate the observed values [<a href="#B6-atmosphere-15-00735" class="html-bibr">6</a>], the solid lines are the LES model values, and the dashed lines are the fitted logarithmic profile, with <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>z</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> = 0.01 m, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>u</mi> </mrow> <mrow> <mo>∗</mo> </mrow> </msub> </mrow> </semantics></math> = 0.34 m/s, and k = 0.41. Note that the H is the shelterbelt canopy height in this case, and logrithmic profile is assumed for domains higher than 1.28 H (&gt;9 m) where the observation was not available.</p>
Full article ">Figure 4
<p>Plots of an instantaneous total wind speed for the xy plane (<b>a</b>) through y = 65 m and xy plane (<b>b</b>) for z = 4 m. Dashed rectangles represent the cross sections of the shelterbelt.</p>
Full article ">Figure 5
<p>Comparison of time-averaged wind profiles (from 500 to 1500 s) at the observation points P<sub>0</sub> to P<sub>5</sub>. The U profiles with coarse mesh (black lines), fine mesh (green lines), and observed U profiles (circles) at the observation points denoted in <a href="#atmosphere-15-00735-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 6
<p>Comparison of time-averaged TKE profiles (from 500 to 1500 s) at the observation points P<sub>0</sub> to P<sub>5</sub>. TKE with coarse mesh (black lines), fine mesh (green lines), and observed TKE (circles) at the observation points denoted in <a href="#atmosphere-15-00735-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 7
<p>Dynamic pressure <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>p</mi> </mrow> <mrow> <mo>′</mo> </mrow> </msup> </mrow> </semantics></math> computed from the model in a vertical cross section at y = 65 m. The value is normalized by mean horizontal momentum. The dashed rectangle represents the x–y cross section of the shelterbelt.</p>
Full article ">Figure 8
<p>Display of average wind speed (<b>left panel</b>, <b>a</b>–<b>d</b>) and corresponding average TKE (<b>right panel</b>, <b>e</b>–<b>h</b>) for the differing vegetation element densities (<span class="html-italic">A</span> values, m<sup>−1</sup>) (<b>a</b>,<b>e</b>) <span class="html-italic">A</span> = 0.19 m<sup>−1</sup>; (<b>b</b>,<b>f</b>) <span class="html-italic">A</span> = 0.75 m<sup>−1</sup>; (<b>c</b>,<b>g</b>) <span class="html-italic">A</span> = 1.50 m<sup>−1</sup>; and (<b>d</b>,<b>h</b>) <span class="html-italic">A</span> = 2.25 m<sup>−1</sup>. The average wind speeds and TKEs are normalized by the wind speed at a point located at x = 70 m and z = 7 m away on the windward side. All shelterbelts in the simulation were 70 m in length, 7 m in height, and 2 m in width. All the incoming winds are perpendicular to the shelterbelts. The dashed rectangle represents the x-y cross section of the shelterbelt.</p>
Full article ">Figure 9
<p>Sensitivity test of shelterbelt width. The shelterbelts x-z cross sections are shown as dashed rectangles. The right panels are averaged wind speeds; the left panels are averaged TKEs. (<b>a</b>,<b>b</b>) the same run with a 2 m shelterbelt width, <span class="html-italic">A</span> = 1.34 m<sup>−1</sup>; (<b>c</b>,<b>d</b>) the same run with 15 m shelterbelt width, <span class="html-italic">A</span> = 1.34 m<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Sensitivity of approaching wind directions with respect to the shelterbelts. All three shelterbelts had the same dimensions (height = 7 m, width = 2 m, length = 70 m) and vegetation element density <span class="html-italic">A</span> = 1.34 m<sup>−1</sup>. All plots are instantaneous (t = 1000 s) wind speed horizontal cross sections at z = 4 m. (<b>a</b>) Instantaneous wind speed with incoming wind perpendicular to the shelterbelt, (<b>b</b>) instantaneous wind speed with incoming wind of 45° with respect to the shelterbelt, (<b>c</b>) instantaneous wind speed with incoming wind of 60° with respect to the shelterbelt, (<b>d</b>) average TKE corresponding to (<b>a</b>,<b>e</b>) average TKE corresponding to (<b>b</b>,<b>f</b>) average TKE corresponding to (<b>c</b>).</p>
Full article ">Figure 11
<p>(<b>a</b>) Domain buildings and forest shelterbelt setup. The northern shelterbelt is 9 m in width and 15 m in height. The eastern shelterbelt is 15 m in width and 15 m in height. (<b>b</b>) Instantaneous horizontal cross section of wind field at z = 12 m at t = 1000 s. The shelterbelts are denoted in dashed red rectangles, and both shelterbelts have a vegetation element density of 1 m<sup>−1</sup>. The wind on the lee side of the shelterbelts is quite different because the northern shelterbelt is narrow than the eastern shelterbelt, and the local wind direction of the eastern shelterbelt is almost parallel with respect to the shelterbelt (see <a href="#sec3dot2dot3-atmosphere-15-00735" class="html-sec">Section 3.2.3</a> for a more detailed explanation).</p>
Full article ">
19 pages, 3490 KiB  
Review
Efficacy and Safety of Intravenous Magnesium Sulfate in Spinal Surgery: A Systematic Review and Meta-Analysis
by Jorge Campos, Jose Luis Bas, Claudia Campos, Gonzalo Mariscal, Teresa Bas and Paloma Bas
J. Clin. Med. 2024, 13(11), 3122; https://doi.org/10.3390/jcm13113122 - 26 May 2024
Viewed by 1327
Abstract
Optimizing pain management in spinal surgery is crucial for preventing adverse events due to delayed mobilization. Magnesium sulfate has potential benefits in spinal surgery because of its analgesic properties and modulation of neurotransmitters and autonomic nervous system. Existing evidence regarding the use of [...] Read more.
Optimizing pain management in spinal surgery is crucial for preventing adverse events due to delayed mobilization. Magnesium sulfate has potential benefits in spinal surgery because of its analgesic properties and modulation of neurotransmitters and autonomic nervous system. Existing evidence regarding the use of magnesium sulfate is partial and controversial, necessitating a comprehensive meta-analysis to evaluate its efficacy and safety. The aim of this study was to conduct a comprehensive meta-analysis to evaluate the efficacy and safety of magnesium sulfate in spinal surgery compared to other available options. This meta-analysis adhered to the PRISMA guidelines. Patients undergoing spinal surgery were included, with the intervention group receiving intravenous magnesium sulfate (MS) at various doses or combinations, whereas the comparison group received other alternatives or a placebo. The efficacy and safety outcomes were assessed. Data were collected from multiple databases and analyzed using Review Manager version 5.4. Heterogeneity was assessed and fixed- or random-effects models were applied. The meta-analysis included eight studies (n = 541). Magnesium sulfate demonstrated significant reductions in pain at 24 h (MD −0.20, 95% CI: −0.39 to −0.02) and opioid consumption (SMD −0.66, 95% CI: −0.95 to −0.38) compared to placebo. Additionally, a decrease in the use of muscle relaxants (SMD −0.91, 95% CI: −1.65 to −0.17) and remifentanil (SMD −1.52, 95% CI: −1.98 to −1.05) was observed. In contrast, an increase in extubation time (MD 2.42, 95% CI: 1.14 to 3.71) and verbal response (MD 1.85, 95% CI: 1.13 to 2.58) was observed compared to dexmedetomidine. In conclusion, magnesium sulfate administration in spinal surgery reduced pain and opioid consumption, and prolonged orientation and verbal response. No significant differences in blood pressure or heart rate were observed between the groups. Full article
(This article belongs to the Special Issue Neurosurgery and Spine Surgery: From Up-to-Date Practitioners)
Show Figures

Figure 1

Figure 1
<p>Flow diagram depicting the study selection process (Preferred Reporting Items for Systematic Reviews and Meta-analyses).</p>
Full article ">Figure 2
<p>Assessment of the risk of bias (green = low risk; red = high risk; white = unknown) [<a href="#B9-jcm-13-03122" class="html-bibr">9</a>,<a href="#B10-jcm-13-03122" class="html-bibr">10</a>,<a href="#B14-jcm-13-03122" class="html-bibr">14</a>,<a href="#B15-jcm-13-03122" class="html-bibr">15</a>,<a href="#B16-jcm-13-03122" class="html-bibr">16</a>,<a href="#B20-jcm-13-03122" class="html-bibr">20</a>,<a href="#B21-jcm-13-03122" class="html-bibr">21</a>,<a href="#B22-jcm-13-03122" class="html-bibr">22</a>].</p>
Full article ">Figure 3
<p>Forest plot displaying pain measured using Visual Analog Scale (VAS) [<a href="#B10-jcm-13-03122" class="html-bibr">10</a>,<a href="#B16-jcm-13-03122" class="html-bibr">16</a>,<a href="#B20-jcm-13-03122" class="html-bibr">20</a>,<a href="#B21-jcm-13-03122" class="html-bibr">21</a>].</p>
Full article ">Figure 4
<p>Forest plot illustrating opioid consumption (<b>a</b>), muscle relaxant consumption (<b>b</b>), remifentanil consumption (<b>c</b>), and vasopressor consumption (<b>d</b>) [<a href="#B9-jcm-13-03122" class="html-bibr">9</a>,<a href="#B10-jcm-13-03122" class="html-bibr">10</a>,<a href="#B14-jcm-13-03122" class="html-bibr">14</a>,<a href="#B16-jcm-13-03122" class="html-bibr">16</a>,<a href="#B20-jcm-13-03122" class="html-bibr">20</a>,<a href="#B21-jcm-13-03122" class="html-bibr">21</a>,<a href="#B22-jcm-13-03122" class="html-bibr">22</a>].</p>
Full article ">Figure 5
<p>Forest plot presenting the results of heart rate at 30 min, 60 min, 120 min, and after extubation [<a href="#B9-jcm-13-03122" class="html-bibr">9</a>,<a href="#B15-jcm-13-03122" class="html-bibr">15</a>,<a href="#B20-jcm-13-03122" class="html-bibr">20</a>,<a href="#B22-jcm-13-03122" class="html-bibr">22</a>].</p>
Full article ">Figure 6
<p>Forest plot showing the mean arterial pressure at 30 min, 60 min, 120 min, and after extubation [<a href="#B9-jcm-13-03122" class="html-bibr">9</a>,<a href="#B10-jcm-13-03122" class="html-bibr">10</a>,<a href="#B15-jcm-13-03122" class="html-bibr">15</a>,<a href="#B20-jcm-13-03122" class="html-bibr">20</a>,<a href="#B22-jcm-13-03122" class="html-bibr">22</a>].</p>
Full article ">Figure 7
<p>Forest plot demonstrating extubation time (<b>a</b>), response to verbal commands (<b>b</b>), and orientation time (<b>c</b>) [<a href="#B10-jcm-13-03122" class="html-bibr">10</a>,<a href="#B15-jcm-13-03122" class="html-bibr">15</a>,<a href="#B16-jcm-13-03122" class="html-bibr">16</a>,<a href="#B21-jcm-13-03122" class="html-bibr">21</a>,<a href="#B22-jcm-13-03122" class="html-bibr">22</a>].</p>
Full article ">
13 pages, 3331 KiB  
Article
Dynamic Thermal Response of Multiple Interface Cracks between a Half-Plane and a Coating Layer under General Transient Temperature Loading
by Mahsa Nourazar, Weilin Yang and Zengtao Chen
Materials 2024, 17(11), 2478; https://doi.org/10.3390/ma17112478 - 21 May 2024
Viewed by 724
Abstract
This paper explores the thermal behavior of multiple interface cracks situated between a half-plane and a thermal coating layer when subjected to transient thermal loading. The temperature distribution is analyzed using the hyperbolic heat conduction theory. In this model, cracks are represented as [...] Read more.
This paper explores the thermal behavior of multiple interface cracks situated between a half-plane and a thermal coating layer when subjected to transient thermal loading. The temperature distribution is analyzed using the hyperbolic heat conduction theory. In this model, cracks are represented as arrays of thermal dislocations, with densities calculated via Fourier and Laplace transformations. The methodology involves determining the temperature gradient within the uncracked region, and these calculations contribute to formulating a singular integral equation specific to the crack problem. This equation is subsequently utilized to ascertain the dislocation densities at the crack surface, which facilitates the estimation of temperature gradient intensity factors for the interface cracks experiencing transient thermal loading. This paper further explores how the relaxation time, loading parameters, and crack dimensions impact the temperature gradient intensity factors. The results can be used in fracture analysis of structures operating at high temperatures and can also assist in the selection and design of coating materials for specific applications, to minimize the damage caused by temperature loading. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic view of a single dislocation at an interface.</p>
Full article ">Figure 2
<p>(<b>a</b>) Temperature variation at the central point on the upper and lower crack faces. (<b>b</b>) Finite element model of the crack problem simulated in Abaqus.</p>
Full article ">Figure 3
<p>(<b>a</b>) Temperature gradient intensity factor for a single crack versus time for different values of <math display="inline"><semantics> <mrow> <mi>η</mi> </mrow> </semantics></math> and <span class="html-italic">h</span>. (<b>b</b>) Temperature gradient intensity factor for a single crack versus time (symmetrical, <span class="html-italic">ηL</span> = 1). (<b>c</b>) Temperature gradient intensity factor for a single crack versus time for different values of the coating’s thermal conductivity (symmetrical, <math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> </mrow> </semantics></math> = 1). (<b>d</b>) Temperature gradient intensity factor for a single crack versus time for different values of the coating’s relaxation time (symmetrical, <math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> </mrow> </semantics></math> = 1).</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) Temperature gradient intensity factor for a single crack versus time for different values of <math display="inline"><semantics> <mrow> <mi>η</mi> </mrow> </semantics></math> and <span class="html-italic">h</span>. (<b>b</b>) Temperature gradient intensity factor for a single crack versus time (symmetrical, <span class="html-italic">ηL</span> = 1). (<b>c</b>) Temperature gradient intensity factor for a single crack versus time for different values of the coating’s thermal conductivity (symmetrical, <math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> </mrow> </semantics></math> = 1). (<b>d</b>) Temperature gradient intensity factor for a single crack versus time for different values of the coating’s relaxation time (symmetrical, <math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> </mrow> </semantics></math> = 1).</p>
Full article ">Figure 4
<p>(<b>a</b>) Temperature gradient intensity factor for two identical cracks versus time (<math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> </mrow> </semantics></math> = 0). (<b>b</b>) Temperature gradient intensity factor for two colinear cracks versus time (<math display="inline"><semantics> <mrow> <mi>η</mi> <mi>L</mi> <mo>(</mo> <mn>1</mn> <mo>)</mo> </mrow> </semantics></math> = 1).</p>
Full article ">Figure 5
<p>(<b>a</b>) Temperature gradient intensity factor for three colinear cracks with different lengths versus time (<span class="html-italic">L</span>(1) = <span class="html-italic">L</span>(2) = <span class="html-italic">L</span>(3) = <span class="html-italic">L</span>). (<b>b</b>) Temperature gradient intensity factor for three colinear cracks with different lengths versus time (<span class="html-italic">L</span>(1) = <span class="html-italic">L</span>(2) = <span class="html-italic">L</span> &lt; <span class="html-italic">L</span>(3)). (<b>c</b>) Temperature gradient intensity factor for three colinear cracks with different lengths versus time (<span class="html-italic">L</span>(1) = <span class="html-italic">L</span>(3) = <span class="html-italic">L</span>, <span class="html-italic">L</span> &lt; <span class="html-italic">L</span>(2)).</p>
Full article ">
24 pages, 75890 KiB  
Article
Coseismic and Early Postseismic Deformation Mechanism Following the 2021 Mw 7.4 Maduo Earthquake: Insights from Satellite Radar Interferometry and GPS
by Chuanzeng Shu, Zhiguo Meng, Qiong Wu, Wei Xiong, Lijia He, Xiaoping Zhang and Dan Xu
Remote Sens. 2024, 16(8), 1399; https://doi.org/10.3390/rs16081399 - 16 Apr 2024
Cited by 1 | Viewed by 981
Abstract
Exploring the deformation mechanism of the 2021 Mw 7.4 Maduo Earthquake is crucial for better understanding the seismic hazard of the faults with low strain rates inside the Bayan Har block. This study leverages deformation information derived from Sentient-1 A/B images and GPS [...] Read more.
Exploring the deformation mechanism of the 2021 Mw 7.4 Maduo Earthquake is crucial for better understanding the seismic hazard of the faults with low strain rates inside the Bayan Har block. This study leverages deformation information derived from Sentient-1 A/B images and GPS data to investigate in detail the co- and postseismic deformation mechanisms using multiple methods. The main results are as follows. First, the postseismic InSAR time series robustly identified the reactivation of the Changmahe fault, indicating the impact of the Maduo event on surrounding active faults. Second, the joint inversion of Interferometric Synthetic Aperture Radar and GPS revealed that (1) there was a complementary and partially overlapping relationship between the coseismic slip and postseismic afterslip of the main rupture; and (2) the Changmahe fault exhibited thrust compression dislocation in the early stage and experienced a sustained compressive effect from afterslip in the one year after the mainshock. Third, modeling the processes of viscoelastic relaxation and poroelastic rebound revealed that the postseismic deformation was probably caused by a combination of afterslip (near-field) and viscoelastic relaxation (near and far field). Fourth, the stress changes driven by the Maduo event revealed that the seismic gaps inside the Maqin-Maqu segment and the Kunlun Pass-Jiangcuo fault will be potential seismic risks in the future. Full article
Show Figures

Figure 1

Figure 1
<p>The seismotectonic background around the Maduo event. The solid red line represents the boundary of blocks in the Tibetan Plateau. The solid black line represents the developed active faults [<a href="#B20-remotesensing-16-01399" class="html-bibr">20</a>,<a href="#B21-remotesensing-16-01399" class="html-bibr">21</a>]. EKLF = East Kunlun fault, KLJF = Kunlun Pass-Jiangcuo fault, DRF = Dari fault, MGF = Maduo-Gande fault, and BYKLF = Bayan Har Mountain Main Peak fault, TDCF = Tibet Dagou-Changmahe fault. (<b>a</b>) Major blocks and developed faults within the Tibetan Plateau. The blue-to-red beach ball sphere in (<b>a</b>) represents large historical earthquakes occurring at the boundary of the Bayan Har block, with color change indicating variations in depth. The black beach ball in (<b>a</b>) represents the Maduo event. The blue arrow indicates the interseismic GPS rate [<a href="#B22-remotesensing-16-01399" class="html-bibr">22</a>,<a href="#B23-remotesensing-16-01399" class="html-bibr">23</a>]. (<b>b</b>) The tectonic background around the epicenter. The white and black dashed boxes indicate the spatial coverage of Sentinel-1A/B data used to obtain co- and postseismic deformation from ascending and descending orbits, respectively. The solid black line box indicates the range of subgraph (<b>c</b>). The blue triangle represents the GPS station that recorded coseismic deformation, and the red triangle represents the near-field GPS station that recorded postseismic deformation. The black beach ball in (<b>b</b>) represents the 1947 M7.75 Dari earthquake [<a href="#B24-remotesensing-16-01399" class="html-bibr">24</a>]. (<b>c</b>) The circle represents the aftershocks after relocation [<a href="#B4-remotesensing-16-01399" class="html-bibr">4</a>], with color change (the color bar in (<b>b</b>)) representing the relative time in relation to the mainshock. The red star shows the relocated epicenter.</p>
Full article ">Figure 2
<p>Coseismic deformation fields of the Maduo event. “Asc” and “Des”, respectively, indicate the ascending and descending orbits of the satellite. (<b>a</b>,<b>b</b>) represent LOS deformation fields from ascending and descending orbits obtained by D-InSAR, respectively. (<b>c</b>,<b>d</b>) represent LOS deformation fields from ascending and descending orbits obtained by POT, respectively. (<b>e</b>,<b>f</b>) represent the azimuth deformation fields from ascending orbit obtained by POT and BOI, respectively. The solid black line indicates the seismogenic rupture trace drawn by coseismic deformation. The red star shows the relocated epicenter.</p>
Full article ">Figure 3
<p>The 3-D coseismic deformation field of the Maduo event. The horizontal deformation vector from the EW (<b>a</b>) and NS (<b>b</b>) deformation components of the 3-D coseismic deformation field is superimposed on the vertical (<b>c</b>) deformation component. Deformation values in the northward, eastward, and upward directions are positive. The red star shows the relocated epicenter. The solid black line indicates the seismogenic rupture trace drawn by coseismic deformation.</p>
Full article ">Figure 4
<p>The postseismic InSAR deformation fields (<b>a</b>–<b>h</b>), the postseismic GPS horizontal displacement (<b>c</b>), and (<b>i</b>–<b>l</b>) the time series of typical deformation points. The dashed black lines AA’–DD’ indicate the profile in <a href="#app1-remotesensing-16-01399" class="html-app">Figure S3a–d</a>. The black boxes labeled as a-d in (<b>c</b>,<b>g</b>) indicate specifically selected points in (<b>i</b>–<b>l</b>). The black dashed curve represents the isopleth with a 5 mm interval, and Figure 9a,b are overlaid as isopleth in (<b>c</b>) and (<b>g</b>), respectively. The black arrow in subgraph (<b>c</b>) indicates the postseismic GPS horizontal displacement in the current period. F6 represents the fault that was reactivated by the Maduo event. The coarse grey curves in (<b>i</b>–<b>l</b>) represent the best-fit curve (see Formula (2)), and the values of the decay coefficient (τ, unit in day) are labeled near the curves. The red star shows the relocated epicenter.</p>
Full article ">Figure 5
<p>The EW displacement component of the postseismic GPS time series. The red dot and the black dot represent the original and corrected GPS time series, respectively. The bluish-green curve is the logarithmic function curve with the optimal fitting (see Formula (1)), and the values of the decay coefficient (τ-τ<sub>e</sub>, τ<sub>n</sub>, unit in day) are labeled near the curves.</p>
Full article ">Figure 6
<p>The NS displacement component of the postseismic GPS time series. The red dot and the black dot represent the original and corrected GPS time series, respectively. The bluish green curve is the logarithmic function curve with the optimal fitting (see Formula (1)), the values of the decay coefficient (τ-τ<sub>e</sub>, τ<sub>n</sub>, unit in day) are labelled near the curves.</p>
Full article ">Figure 7
<p>The 2.5-D postseismic deformation fields (<b>a</b>–<b>h</b>). The deformation values in the quasi-E and the quasi-upward directions are positive. The black dashed lines AA′–DD′ indicate the profiles of <a href="#app1-remotesensing-16-01399" class="html-app">Figure S3e–h</a>. The black dashed line E–E′ indicates the profile of Figure 10b. The red star shows the relocated epicenter.</p>
Full article ">Figure 8
<p>The co- (<b>a</b>,<b>b</b>) postseismic (1 year) slip distribution of the Maduo event. The white dashed line denotes the coseismic slip isopleth at 1m intervals. The black circle represents the relocated aftershocks distribution. The change of geodetic moment release with depth of co- (<b>c</b>,<b>d</b>) postseismic slip. The red dots indicate the seismic moment corresponding to each depth, and the histogram indicates the slope change of the seismic moment with depth. (<b>e</b>,<b>f</b>) indicate the trade-off curve between model roughness and normalized misfit of the co- and postseismic (1 year) slip distribution, respectively. The red dots in (<b>e</b>,<b>f</b>) represent the preferred smoothing factor. The red star shows the relocated epicenter.</p>
Full article ">Figure 9
<p>The simulated LOS deformation from ascending and descending orbits and GPS horizontal displacements ~1 year after the mainshock based on the models of VR (<b>a</b>,<b>b</b>) and PR (<b>c</b>,<b>d</b>). The red triangles indicate the near-field GPS stations. The red star shows the relocated epicenter.</p>
Full article ">Figure 10
<p>The tectonic background surrounding fault F6 (<b>a</b>), the deformation profile across fault F6 (<b>b</b>), the schematic diagram illustrating the reactivation mechanism of fault F6 (<b>c</b>), and the slip distribution of the fault F6 during the early stage (<b>d</b>) and the one year (<b>e</b>) after the mainshock. The position of the E–E’ profile is indicated in <a href="#remotesensing-16-01399-f007" class="html-fig">Figure 7</a>f. (<b>f</b>) illustrates the trade-off curve between model roughness and normalized misfit of the slip distribution in the early stage after the mainshock. EKLF = East Kunlun fault, KLJF = Kunlun Pass-Jiangcuo fault, DRF = Dari fault, MGF = Maduo-Gande fault, and TDCF = Tibet Dagou-Changmahe fault. The red star shows the relocated epicenter.</p>
Full article ">Figure 11
<p>The coseismic dCFS of the Maduo event. The thick red line shows the Maqin-Maqu segment. The lake filled with white color represents Donggei-Cuona Lake. EKLF = East Kunlun fault, KLJF = Kunlun Pass-Jiangcuo fault, DRF = Dari fault, MGF = Maduo-Gande fault, and TDCF = Tibet Dagou-Changmahe fault. The red star shows the relocated epicenter.</p>
Full article ">
18 pages, 3679 KiB  
Article
Breathing Right… or Left! The Effects of Unilateral Nostril Breathing on Psychological and Cognitive Wellbeing: A Pilot Study
by Maria Elide Vanutelli, Chiara Grigis and Claudio Lucchiari
Brain Sci. 2024, 14(4), 302; https://doi.org/10.3390/brainsci14040302 - 23 Mar 2024
Viewed by 4817
Abstract
The impact of controlled breathing on cognitive and affective processing has been recognized since ancient times, giving rise to multiple practices aimed at achieving different psychophysical states, mostly related to mental clarity and focus, stress reduction, and relaxation. Previous scientific research explored the [...] Read more.
The impact of controlled breathing on cognitive and affective processing has been recognized since ancient times, giving rise to multiple practices aimed at achieving different psychophysical states, mostly related to mental clarity and focus, stress reduction, and relaxation. Previous scientific research explored the effects of forced unilateral nostril breathing (UNB) on brain activity and emotional and cognitive functions. Some evidence concluded that it had a contralateral effect, while other studies presented controversial results, making it difficult to come to an unambiguous interpretation. Also, a few studies specifically addressed wellbeing. In the present study, we invited a pilot sample of 20 participants to take part in an 8-day training program for breathing, and each person was assigned to either a unilateral right nostril (URNB) or left nostril breathing condition (ULNB). Then, each day, we assessed the participants’ wellbeing indices using their moods and mind wandering scales. The results revealed that, after the daily practice, both groups reported improved wellbeing perception. However, the effect was specifically related to the nostril involved. URNB produced more benefits in terms of stress reduction and relaxation, while ULNB significantly and increasingly reduced mind-wandering occurrences over time. Our results suggest that UNB can be effectively used to increase wellbeing in the general population. Additionally, they support the idea that understanding the effects of unilateral breathing on wellbeing and cognition requires a complex interpretive model with multiple brain networks to address bottom-up and top-down processes. Full article
(This article belongs to the Section Neuropsychology)
Show Figures

Figure 1

Figure 1
<p>An outline of the experimental procedure. The participants were first briefed, and the inclusion/exclusion criteria were checked (I). Afterward, they read and signed the informed consent forms, provided demographical information and handedness and filled out personality questionnaires (II). Then, the training began (day 1, III). First, they were asked to complete the mood scales. Second, they executed the breathing technique. Third, they completed the mood scales again. Fourth, they completed the MEWS for intrusive occurrences. From day 2 to day 7 (IV), the participants continued with solo training following the same procedure. Finally, (V), the participants concluded the training with the last in-person session, as in phase III.</p>
Full article ">Figure 2
<p>Line charts of mean self-assessed scores for stressed and restless (<b>up</b>), calmness (<b>center</b>), as well as clear minded and happy moods (<b>bottom</b>) for each day before (blue) and after (orange) breathing practice.</p>
Full article ">Figure 3
<p>After each daily practice, the participants assessed their experiences with mind-wandering intrusions. In the plot above, the post-practice MEWS scores are represented for each day.</p>
Full article ">Figure 4
<p>Before and after each daily practice, the participants completed the mood scales about stressed, restless, calm, happy, and clear minded states. In the histogram above, the mean mood states before (blue) and after (orange) the daily practice are represented. Over each comparison, significant and non-significant (n.s.) effects are indicated.</p>
Full article ">Figure 5
<p>Mean mind wandering scores as recorded by participants on day 1 (light green) and day 8 (dark green) of training.</p>
Full article ">Figure 6
<p>Pairwise comparisons of daily mind wandering values with Friedman test. Significant values were adjusted by Bonferroni correction for multiple comparisons. Each node represents daily assessment and shows average rank. Significant comparisons are displayed in Bordeaux; non-significant comparisons are displayed in blue.</p>
Full article ">Figure 7
<p>Mean delta values (improvement after practice) for mood states, self-assessed by URNB (yellow) and ULNB (water green) groups.</p>
Full article ">Figure 8
<p>Pairwise comparisons of daily mind wandering values with Friedman test for URNB group (<b>left</b>) and ULNB group (<b>right</b>). Significance values were adjusted by Bonferroni correction for multiple comparisons. Each node represents daily assessment and shows average rank. Significant comparisons are displayed in Bordeaux; non-significant comparisons are displayed in blue.</p>
Full article ">Figure 9
<p>Line charts of mean MEWS scores for each day for URNB (blue) and ULNB (orange) groups.</p>
Full article ">
15 pages, 2094 KiB  
Article
Optimal Design of Group Orthogonal Phase-Coded Waveforms for MIMO Radar
by Tianqu Liu, Jinping Sun, Guohua Wang, Xianxun Yao and Yaqiong Qiao
Mathematics 2024, 12(6), 903; https://doi.org/10.3390/math12060903 - 19 Mar 2024
Viewed by 1172
Abstract
Digital radio frequency memory (DRFM) has emerged as an advanced technique to achieve a range of jamming signals, due to its capability to intercept waveforms within a short time. multiple-input multiple-output (MIMO) radars can transmit agile orthogonal waveform sets for different pulses to [...] Read more.
Digital radio frequency memory (DRFM) has emerged as an advanced technique to achieve a range of jamming signals, due to its capability to intercept waveforms within a short time. multiple-input multiple-output (MIMO) radars can transmit agile orthogonal waveform sets for different pulses to combat DRFM-based jamming, where any two groups of waveform sets are also orthogonal. In this article, a group orthogonal waveform optimal design model is formulated in order to combat DRFM-based jamming by flexibly designing waveforms for MIMO radars. Aiming at balancing the intra- and intergroup orthogonal performances, the objective function is defined as the weighted sum of the intra- and intergroup orthogonal performance metrics. To solve the formulated model, in this article, a group orthogonal waveform design algorithm is proposed. Based on a primal-dual-type method and proper relaxations, the proposed algorithm transforms the original problem into a series of simple subproblems. Numerical results demonstrate that the obtained group orthogonal waveforms have the ability to flexibly suppress DRFM-based deceptive jamming, which is not achievable using p-majorization–minimization (p-MM) and primal-dual, two of the most advanced orthogonal waveform design algorithms. Full article
Show Figures

Figure 1

Figure 1
<p>The convergence curves of <span class="html-italic">γ</span> and <span class="html-italic">ε<sub>g</sub></span>, <span class="html-italic">g =</span> 1,2,…,<span class="html-italic">G</span>. (<b>a</b>) <span class="html-italic">w =</span> 0.1, <span class="html-italic">M =</span> 2, <span class="html-italic">G =</span> 2, <span class="html-italic">N =</span> 256; (<b>b</b>) <span class="html-italic">w =</span> 0.7, <span class="html-italic">M =</span> 8, <span class="html-italic">G =</span> 8, <span class="html-italic">N =</span> 256.</p>
Full article ">Figure 2
<p>The convergence curves of <span class="html-italic">γ</span> and <span class="html-italic">ε<sub>g</sub></span>, <span class="html-italic">g =</span> 1,2,…,<span class="html-italic">G</span>. (<b>a</b>) <span class="html-italic">M =</span> 2, <span class="html-italic">G =</span> 2, <span class="html-italic">N =</span> 256; (<b>b</b>) <span class="html-italic">M =</span> 8, <span class="html-italic">G =</span> 8, <span class="html-italic">N =</span> 256.</p>
Full article ">Figure 3
<p>The correlation peak value obtained by the proposed algorithm when <span class="html-italic">M =</span> 2, <span class="html-italic">G =</span> 2, and <span class="html-italic">N =</span> 256: (<b>a</b>) autocorrelation; (<b>b</b>) intragroup cross-correlation; (<b>c</b>) intergroup cross-correlation, and when <span class="html-italic">M =</span> 8, <span class="html-italic">G =</span> 8, and <span class="html-italic">N =</span> 256: (<b>d</b>) autocorrelation; (<b>e</b>) intragroup cross-correlation; (<b>f</b>) intergroup cross-correlation.</p>
Full article ">Figure 4
<p>The effect of parameters <span class="html-italic">M</span>, <span class="html-italic">N</span>, and <span class="html-italic">G</span> on the obtained metric values. (<b>a</b>) <span class="html-italic">w =</span> 0.3, <span class="html-italic">M =</span> 2, <span class="html-italic">G =</span> 2; (<b>b</b>) <span class="html-italic">w =</span> 0.875, <span class="html-italic">M =</span> 2, <span class="html-italic">G =</span> 2; (<b>c</b>) <span class="html-italic">w =</span> 0.5, <span class="html-italic">M =</span> 8, <span class="html-italic">N =</span> 256; (<b>d</b>) <span class="html-italic">w =</span> 0.5, <span class="html-italic">G =</span> 8, <span class="html-italic">N =</span> 256.</p>
Full article ">Figure 5
<p>The block diagram of the matched filtering and digital beam forming processes.</p>
Full article ">Figure 6
<p>Angle–range images formed with MIMO radar using fixed and agile waveforms under DRFM-based deceptive jamming. (<b>a</b>) Fixed waveform; (<b>b</b>) primal-dual; (<b>c</b>) Multi-CAN; (<b>d</b>) <span class="html-italic">p</span>-MM; (<b>e</b>) <span class="html-italic">w =</span> 0.9, <span class="html-italic">M =</span> 3, <span class="html-italic">G =</span> 2, <span class="html-italic">N =</span> 256; (<b>f</b>) <span class="html-italic">w =</span> 0.1, <span class="html-italic">M =</span> 3, <span class="html-italic">G =</span> 2, <span class="html-italic">N =</span> 256.</p>
Full article ">
21 pages, 7174 KiB  
Article
An Informed-Bi-Quick RRT* Algorithm Based on Offline Sampling: Motion Planning Considering Multiple Constraints for a Dual-Arm Cooperative System
by Qinglei Zhang, Yunfeng Liu, Jiyun Qin and Jianguo Duan
Actuators 2024, 13(2), 75; https://doi.org/10.3390/act13020075 - 14 Feb 2024
Cited by 1 | Viewed by 1810
Abstract
Aiming to address problems such as low sampling success rate and long computation time in the motion planning of a dual-arm cooperative system with multiple constraints, this paper proposes an Informed-Bi-Quick RRT* algorithm based on offline sampling. First, in the process of pre-sampling, [...] Read more.
Aiming to address problems such as low sampling success rate and long computation time in the motion planning of a dual-arm cooperative system with multiple constraints, this paper proposes an Informed-Bi-Quick RRT* algorithm based on offline sampling. First, in the process of pre-sampling, the new algorithm relaxes the approximation of constrained manifolds by introducing the idea of incremental construction, and it incorporates the stochastic gradient descent method to replace global random sampling with local random sampling, which enriches the data set and shortens the offline sampling time of the data set. Second, the new algorithm improves the original Quick-RRT* algorithm by combining the two-tree idea and the multi-target bias expansion strategy, and it improves the adaptability of the algorithm to different obstacle environments. In addition, the loosely constrained motion and tightly constrained motion in the two-arm cooperative system are analyzed, and the adaptive planning of the two-arm trajectory in different motions is described in detail. In this paper, the two-arm cooperative model constructed with UR5 and UR10 robot arms is studied, and the ability of the proposed algorithm to deal with multiple constraints is verified by simulating assembly and handling tasks. The experimental results show that compared with other methods, the proposed algorithm has obvious advantages in path quality and planning efficiency. Full article
(This article belongs to the Section Control Systems)
Show Figures

Figure 1

Figure 1
<p>Flow diagram of off-line data set construction.</p>
Full article ">Figure 2
<p>Axle hole assembly relative pose diagram.</p>
Full article ">Figure 3
<p>Handling relative pose diagram.</p>
Full article ">Figure 4
<p>Algorithm comparison in a simple environment.</p>
Full article ">Figure 5
<p>Algorithm comparison in a complex environment.</p>
Full article ">Figure 6
<p>Real environment of the two-arm collaboration system.</p>
Full article ">Figure 7
<p>Initial pose of two−arm collaboration system.</p>
Full article ">Figure 8
<p>Cooperative system’s position of assembly task.</p>
Full article ">Figure 9
<p>Z−axis position of UR5 in assembly task.</p>
Full article ">Figure 10
<p>Z−axis position of UR10 in assembly task.</p>
Full article ">Figure 11
<p>Cooperative system’s position of handling task.</p>
Full article ">Figure 12
<p>Z-axis position of UR5 in handling task.</p>
Full article ">Figure 13
<p>Z-axis position of UR10 in handling task.</p>
Full article ">Figure 14
<p>Cooperative system’s position.</p>
Full article ">
Back to TopTop