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 (124)

Search Parameters:
Keywords = lattice path

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 3028 KiB  
Article
Numerical Solutions to the Variational Problems by Dijkstra’s Path-Finding Algorithm
by Thanaporn Arunthong, Laddawan Rianthakool, Khanchai Prasanai, Chakrit Na Takuathung, Sakchai Chomkokard, Wiwat Wongkokua and Noparit Jinuntuya
Appl. Sci. 2024, 14(22), 10674; https://doi.org/10.3390/app142210674 - 19 Nov 2024
Viewed by 293
Abstract
In this work, we propose the general idea of using a path-finding algorithm to solve a variational problem. By interpreting a variational problem of finding the function that minimizes a functional integral as a shortest path finding, we can apply the shortest path-finding [...] Read more.
In this work, we propose the general idea of using a path-finding algorithm to solve a variational problem. By interpreting a variational problem of finding the function that minimizes a functional integral as a shortest path finding, we can apply the shortest path-finding algorithm to numerically estimate the optimal function. This can be achieved by discretizing the continuous domain of the variational problem into a spatially weighted graph. The weight of each edge is defined according to the function of the original problem. We adopt the Moser lattice as the discretization scheme since it provides adjustable connections around a vertex. We find that this number of connections is crucial to the estimation of an accurate optimal path. Dijkstra’s shortest path-finding algorithm was chosen due to its simplicity and convenience in implementation. We validate our proposal by applying Dijkstra’s path-finding algorithm to numerically solve three famous variational problems, i.e., the optical ray tracing, the brachistochrone, and the catenary problems. The first two are examples of problems with no constraint. The standard Dijkstra’s algorithm can be directly applied. The third problem is an example of a problem with an isoperimetric constraint. We apply the Lagrangian relaxation technique to relax the optimization in the standard Dijkstra algorithm to incorporate the constraint. In all cases, when the number of sublattices is large enough, the results agree well with the analytic solutions. In all cases, the same path-finding code is used, regardless of the problem details. Our approaches provide more insight and promise to be more flexible than conventional numerical methods. We expect that our method can be useful in practice when an investigation of the optimal path in a complex problem is needed. Full article
Show Figures

Figure 1

Figure 1
<p>The shortest path from a to b, with the smallest accumulated weight of 20.</p>
Full article ">Figure 2
<p>An example of a 3 × 3 Moser lattice (large dots) with 5 sub lattices (small dots).</p>
Full article ">Figure 3
<p>Examples of paths join vertices a and b.</p>
Full article ">Figure 4
<p>An example of the designation of the weight of an edge connecting the vertices a and b is the ToF t<sub>cd</sub>. The purple arrows are the velocity vector of the wave at vertices c and d, with wave speeds v<sub>c</sub> and v<sub>d</sub>, respectively. The weight is defined as the ToF from vertices c to d along the straight edge, which can be estimated as the ratio of the Euclidian distance s<sub>cd</sub> and the average speed (v<sub>c</sub> + v<sub>d</sub>)/2. The weights of the other edges can be defined similarly.</p>
Full article ">Figure 5
<p>Simulation of wave refraction through two homogeneous mediums.</p>
Full article ">Figure 6
<p>Simulation of the onset of total internal reflection.</p>
Full article ">Figure 7
<p>Comparison of the ray tracing in the axial index of refraction profile. Solid lines are the exact solutions, and the dashed lines are the simulation results from Dijkstra’s algorithm.</p>
Full article ">Figure 8
<p>Comparison of the ray tracing in the axial index of refraction profile. Solid lines are the exact solutions, and the dashed lines are the simulation results from Dijkstra’s algorithm.</p>
Full article ">Figure 9
<p>Comparison of the catenary curves for various lengths. Solid lines are the exact solutions, and the dashed lines are the simulation results from Dijkstra’s algorithm.</p>
Full article ">
14 pages, 19855 KiB  
Article
Effect of Deformed Prior Austenite Characteristics on Reverse Phase Transformation and Deformation Behavior of High-Strength Medium-Mn Steel
by Ying Dong, Jingwen Zhang, Tao Liu, Mingxing Ma, Lei Zhu, Chengjun Zhu and Linxiu Du
Materials 2024, 17(22), 5618; https://doi.org/10.3390/ma17225618 - 17 Nov 2024
Viewed by 544
Abstract
In this study, microstructure evolution during prior austenite decomposition and reverse phase transformation processes was revealed in a high-strength medium-Mn steel. Furthermore, the relationship between deformed prior austenite characteristics and deformation behavior was studied. The results indicated that the recovery and recrystallization of [...] Read more.
In this study, microstructure evolution during prior austenite decomposition and reverse phase transformation processes was revealed in a high-strength medium-Mn steel. Furthermore, the relationship between deformed prior austenite characteristics and deformation behavior was studied. The results indicated that the recovery and recrystallization of the deformed prior austenite were significantly inhibited during hot rolling in the non-recrystallized zone, the grain size was obviously refined along the normal direction (ND), and that the strain hardening of prior austenite via hot deformation could increase the resistance of shear transformation, resulting in the preservation of high-density lattice defects in the quenched martensite matrix. Before the nucleation of intercritical austenite, the dislocation and grain boundary can provide fast diffusion paths for C and Mn, and the enrichment of C and Mn before intercritical austenite formation can reduce the critical temperature of ferrite/austenite transformation. The nucleated sites and driving force for intercritical austenite were strongly increased by rolling in the non-recrystallization region. The resistance of crack propagation was found to be enhanced by the sustained transformation-induced plasticity (TRIP) effect (via retained austenite with different stability) and for the laminated microstructure, the optimum properties were obtained as being a combination of yield strength of 748 MPa, tensile strength of 952 MPa, and total elongation of 26.2%. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustrations of thermomechanical controlled processes and heat treatment processes and initial conditions of diffusion simulation.</p>
Full article ">Figure 2
<p>Micrograph of water-quenched experimental steels and hardness. (<b>a</b>) optical microscope (OM) micrograph of R820 sample; (<b>b</b>) OM micrograph of R1000 sample; (<b>c</b>) transmission electron microscope micrograph of R1000 sample.</p>
Full article ">Figure 3
<p>Scanning electron microscope (SEM) micrograph of experimental steels and equilibrium phase diagram. (<b>a</b>,<b>d</b>) L570 sample; (<b>b</b>,<b>e</b>) H570 sample; (<b>c</b>) equilibrium phase diagram; (<b>f</b>) L630 sample.</p>
Full article ">Figure 4
<p>SEM micrograph, Mn distribution, and the corresponding equilibrium calculation of experimental steels. (<b>a</b>) SEM micrograph of L630 sample; (<b>d</b>) SEM micrograph of H630 sample, the prior austenite grain boundaries were marked with the yellow dotted lines; (<b>b</b>,<b>e</b>) the corresponding Mn distribution; (<b>c</b>) Mn and C content in FCC; (<b>f</b>) Mn and C content in BCC.</p>
Full article ">Figure 5
<p>Microstructure of H630 sample analyzed by electron backscattered diffractometer (EBSD). (<b>a</b>) Combined map of band contrast image and phase image; (<b>b</b>) Inverse pole figure; (<b>c</b>,<b>d</b>) the enlargement of the corresponding squares region in (<b>a</b>,<b>b</b>); (<b>e</b>) crystallographic analysis of several grains, which are pointed out as a, b, c, d, e, and f in <a href="#materials-17-05618-f006" class="html-fig">Figure 6</a>c, the arrows represent the specific orientation.</p>
Full article ">Figure 6
<p>DICTRA calculation results of austenite/ferrite interface migration and element diffusion during intercritical austenite growth process. (<b>a</b>,<b>b</b>) Phase interface position versus isothermal duration; (<b>c</b>,<b>d</b>) Mn/C content profiles after isothermal holding at different temperatures for 0.5 h; (<b>e</b>,<b>f</b>) Mn/C content profiles after isothermal holding at 650 for different duration.</p>
Full article ">Figure 7
<p>X-ray diffractometer patterns of experimental steels. (<b>a</b>) L570, L610, and L630 samples; (<b>b</b>) H570, H610, and H630 samples.</p>
Full article ">Figure 8
<p>Engineering stress–strain curves, true strain–stress (TT), and true strain–work hardening rate (TW) curves of experimental steels. (<b>a</b>) Engineering stress-strain curves of samples with flat–elongated PAGs; (<b>b</b>) combined TW and TT curve of L630 sample, the letters a–f represent the sudden change points; (<b>c</b>) combined TT and TW curves of samples L570 and L630; (<b>d</b>) engineering stress-strain curves of samples with equiaxed-recrystallized PAGs; (<b>e</b>) TW of samples with equiaxed-recrystallized PAGs; (<b>f</b>) TW curves of samples L630 and H630.</p>
Full article ">Figure 9
<p>Fracture morphology of tensile samples after being subjected to different processes. Zone A represents the rapid crack propagation zone, zone B represents the plastic deformation concentration zone. (<b>a</b>) L570; (<b>b</b>) L630; (<b>c</b>) H570; (<b>d</b>) H630; (<b>a-1</b>,<b>b-1</b>,<b>c-1</b>,<b>d-1</b>) zone A; (<b>a-2</b>,<b>b-2</b>,<b>c-2</b>,<b>d-2</b>) zone B.</p>
Full article ">
9 pages, 3248 KiB  
Article
Crack Control in Additive Manufacturing by Leveraging Process Parameters and Lattice Design
by Jun Hak Lee, Seong Je Park, Jeongho Yang, Seung Ki Moon and Jiyong Park
Micromachines 2024, 15(11), 1361; https://doi.org/10.3390/mi15111361 - 10 Nov 2024
Viewed by 589
Abstract
This study investigates the design of additive manufacturing for controlled crack propagation using process parameters and lattice structures. We examine two lattice types—octet-truss (OT) and diamond (DM)—fabricated via powder bed fusion with Ti-6Al-4V. Lattice structures are designed with varying densities (10%, 30%, and [...] Read more.
This study investigates the design of additive manufacturing for controlled crack propagation using process parameters and lattice structures. We examine two lattice types—octet-truss (OT) and diamond (DM)—fabricated via powder bed fusion with Ti-6Al-4V. Lattice structures are designed with varying densities (10%, 30%, and 50%) and process using two different laser energies. Using additive-manufactured specimens, Charpy impact tests are conducted to evaluate the fracture behavior and impact energy levels of the specimens. Results show that the type of the lattice structures, the density of the lattice structures, and laser energy significantly influence crack propagation patterns and impact energy. OT exhibits straighter crack paths, while DM demonstrates more random fracture patterns. Higher-density lattices and increased laser energy generally improve the impact energy. DM consistently outperformed OT in the impact energy for angle specimens, while OT showed superior performance in stair specimens. Finally, a case study demonstrates the potential for combining OT and DM structures to guide crack propagation along predetermined paths, offering a novel approach to protect critical components during product failure. Full article
(This article belongs to the Special Issue Laser Additive Manufacturing of Metallic Materials, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The lattice structure for (<b>a</b>) OT and (<b>b</b>) DM with unit cell size.</p>
Full article ">Figure 2
<p>Dimensions of Charpy impact specimens with lattice structures.</p>
Full article ">Figure 3
<p>Geometry information of specimens for crack guidance: (<b>a</b>) angle and (<b>b</b>) stair.</p>
Full article ">Figure 4
<p>(<b>a</b>) Fracture pattern according to the lattice structures and laser energies. The impact energy of (<b>b</b>) OT and (<b>c</b>) DM according to the laser energies and density of lattice structures.</p>
Full article ">Figure 5
<p>Crack propagation pattern and impact energy of angle specimens with (<b>a</b>) OT and (<b>b</b>) DM in two laser energies.</p>
Full article ">Figure 6
<p>Crack propagation pattern and impact energy of stair specimens with (<b>a</b>) OT and (<b>b</b>) DM in two laser energies. (Red box: cases in which a crack does not propagate along the lattice structures.).</p>
Full article ">Figure 7
<p>(<b>a</b>) Design of a part for a case study. (<b>b</b>) Before and after impact test of designed part.</p>
Full article ">
62 pages, 9349 KiB  
Article
Fokker-Planck Central Moment Lattice Boltzmann Method for Effective Simulations of Fluid Dynamics
by William Schupbach and Kannan Premnath
Fluids 2024, 9(11), 255; https://doi.org/10.3390/fluids9110255 - 29 Oct 2024
Viewed by 623
Abstract
We present a new formulation of the central moment lattice Boltzmann (LB) method based on a minimal continuous Fokker-Planck (FP) kinetic model, originally proposed for stochastic diffusive-drift processes (e.g., Brownian dynamics), by adapting it as a collision model for the continuous Boltzmann equation [...] Read more.
We present a new formulation of the central moment lattice Boltzmann (LB) method based on a minimal continuous Fokker-Planck (FP) kinetic model, originally proposed for stochastic diffusive-drift processes (e.g., Brownian dynamics), by adapting it as a collision model for the continuous Boltzmann equation (CBE) for fluid dynamics. The FP collision model has several desirable properties, including its ability to preserve the quadratic nonlinearity of the CBE, unlike that based on the common Bhatnagar-Gross-Krook model. Rather than using an equivalent Langevin equation as a proxy, we construct our approach by directly matching the changes in different discrete central moments independently supported by the lattice under collision to those given by the CBE under the FP-guided collision model. This can be interpreted as a new path for the collision process in terms of the relaxation of the various central moments to “equilibria”, which we term as the Markovian central moment attractors that depend on the products of the adjacent lower order moments and a diffusion coefficient tensor, thereby involving of a chain of attractors; effectively, the latter are nonlinear functions of not only the hydrodynamic variables, but also the non-conserved moments; the relaxation rates are based on scaling the drift coefficient by the order of the moment involved. The construction of the method in terms of the relevant central moments rather than via the drift and diffusion of the distribution functions directly in the velocity space facilitates its numerical implementation and analysis. We show its consistency to the Navier-Stokes equations via a Chapman-Enskog analysis and elucidate the choice of the diffusion coefficient based on the second order moments in accurately representing flows at relatively low viscosities or high Reynolds numbers. We will demonstrate the accuracy and robustness of our new central moment FP-LB formulation, termed as the FPC-LBM, using the D3Q27 lattice for simulations of a variety of flows, including wall-bounded turbulent flows. We show that the FPC-LBM is more stable than other existing LB schemes based on central moments, while avoiding numerical hyperviscosity effects in flow simulations at relatively very low physical fluid viscosities through a refinement to a model founded on kinetic theory. Full article
(This article belongs to the Special Issue Lattice Boltzmann Methods: Fundamentals and Applications)
Show Figures

Figure 1

Figure 1
<p>Routes for the derivation of the Boltzmann equation and the modeling of its collision term via BGK or FP approach under appropriate approximations, and their applications to representing the dynamics in fluids and plasmas (inspired from [<xref ref-type="bibr" rid="B41-fluids-09-00255">41</xref>]).</p>
Full article ">Figure 2
<p>Representation of collision processes at different levels of modeling description and the associated mathematical nature of the collision operator.</p>
Full article ">Figure 3
<p>Streamlines for two-dimensional lid-driven square cavity flow computed using the FPC-LBM at Reynolds numbers of <inline-formula><mml:math id="mm866"><mml:semantics><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> = 1000, 3200, 5000, and 7500. The formation of secondary and tertiary vortices is consistent with those in the benchmark results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>] for each Reynolds number shown here.</p>
Full article ">Figure 4
<p>Comparisons of the horizontal velocity component along the vertical centerline in a two-dimensional lid-driven square cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>]. (<bold>a</bold>) <inline-formula><mml:math id="mm867"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>b</bold>) <inline-formula><mml:math id="mm868"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>3200</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>) <inline-formula><mml:math id="mm869"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>5000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>d</bold>) <inline-formula><mml:math id="mm870"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>7500</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 5
<p>Comparisons of the vertical velocity component along the horizontal centerline in a two-dimensional lid-driven square cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ghia et al. (1982) [<xref ref-type="bibr" rid="B69-fluids-09-00255">69</xref>]. (<bold>a</bold>) <inline-formula><mml:math id="mm871"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>b</bold>) <inline-formula><mml:math id="mm872"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>3200</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>) <inline-formula><mml:math id="mm873"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>5000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>d</bold>) <inline-formula><mml:math id="mm874"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>7500</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 6
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm875"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm876"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm877"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm878"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm879"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm880"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm881"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM. The MCM-LBM is seen to sufficiently capture the physics of this case for all of the grid resolutions and Mach numbers considered here as it has not caused the formation of any spurious secondary vortices.</p>
Full article ">Figure 7
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm882"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm883"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm884"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm885"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm886"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm887"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm888"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Fokker-Planck equilibria based FPC-LBM. The FPC-LBM is seen to sufficiently capture the physics of this case for all of the grid resolutions and Mach numbers considered here as it has not caused the formation of any spurious secondary vortices. The FPC-LBM and MCM-LBM results are almost indistinguishable from one another for these cases indicating the robustness of central moment collision models in general and that we must consider more extreme cases to see significant differences between them.</p>
Full article ">Figure 8
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm889"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of higher Mach numbers of <inline-formula><mml:math id="mm890"><mml:semantics><mml:mrow><mml:mn>0.4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm891"><mml:semantics><mml:mrow><mml:mn>0.5</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm892"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm893"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm894"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm895"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM. It is seen that at larger Mach numbers and coarse grid resolutions, the MCM-LBM becomes unstable with the formation of spurious secondary vortices which form on the layers for the cases of <inline-formula><mml:math id="mm896"><mml:semantics><mml:mrow><mml:mrow><mml:mi>M</mml:mi><mml:mi>a</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>0.4</mml:mn><mml:mo>,</mml:mo><mml:mn>0.5</mml:mn><mml:mo>,</mml:mo></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm897"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm898"><mml:semantics><mml:mrow><mml:msup><mml:mi>L</mml:mi><mml:mn>2</mml:mn></mml:msup><mml:mo>=</mml:mo><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 9
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm899"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of higher Mach numbers of <inline-formula><mml:math id="mm900"><mml:semantics><mml:mrow><mml:mn>0.4</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm901"><mml:semantics><mml:mrow><mml:mn>0.5</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm902"><mml:semantics><mml:mrow><mml:mn>0.6</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm903"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm904"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm905"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) computed using the Maxwellian equilibria based MCM-LBM.Vorticity contours at <inline-formula><mml:math id="mm906"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>. It is seen that the FPC-LBM remains stable and no spurious vortices are formed for the same cases of higher Mach numbers and grid resolutions as those shown previously for the MCM-LBM which did not remain stable (see <xref ref-type="fig" rid="fluids-09-00255-f008">Figure 8</xref>). This indicates that the FPC-LBM is a more robust collision model when compared to the MCM-LBM for simulations at coarse grid resolutions and at relatively large Mach numbers.</p>
Full article ">Figure 10
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm907"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm908"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm909"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm910"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm911"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm912"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm913"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) at an extremely large bulk viscosity by setting the relaxation parameter associated with the bulk viscosity to <inline-formula><mml:math id="mm914"><mml:semantics><mml:mrow><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.35</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and computed using the Maxwellian equilibria based MCM-LBM. It is seen that MCM-LBM becomes unstable under an extreme increase in bulk viscosity, especially for coarse grid resolutions which become progressively worse as the Mach number is increased. By increasing the relaxation time <inline-formula><mml:math id="mm915"><mml:semantics><mml:mrow><mml:msub><mml:mi>τ</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>1</mml:mn><mml:mo>/</mml:mo><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub></mml:mrow></mml:semantics></mml:math></inline-formula> associated with the bulk viscosity to <inline-formula><mml:math id="mm916"><mml:semantics><mml:mrow><mml:mn>2.857</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> from being <inline-formula><mml:math id="mm917"><mml:semantics><mml:mrow><mml:mo>∼</mml:mo><mml:mn>1.0</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> or smaller, the range of beneficial limits is exceeded, and instead the simulations begin to numerically destabilize with MCM-LBM.</p>
Full article ">Figure 11
<p>Vorticity contours of doubly periodic shear layers that roll up due to an applied perturbation at <inline-formula><mml:math id="mm918"><mml:semantics><mml:mrow><mml:mi>t</mml:mi><mml:mo>=</mml:mo><mml:mn>1</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> for different sets of lower Mach numbers of <inline-formula><mml:math id="mm919"><mml:semantics><mml:mrow><mml:mn>0.05</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm920"><mml:semantics><mml:mrow><mml:mn>0.2</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm921"><mml:semantics><mml:mrow><mml:mn>0.3</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> (along rows) and at grid resolutions of <inline-formula><mml:math id="mm92211"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm923"><mml:semantics><mml:msup><mml:mn>128</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, and <inline-formula><mml:math id="mm924"><mml:semantics><mml:msup><mml:mn>256</mml:mn><mml:mn>2</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> (along columns) at an extremely large bulk viscosity by setting the relaxation parameter associated with the bulk viscosity to <inline-formula><mml:math id="mm925"><mml:semantics><mml:mrow><mml:msub><mml:mi>ω</mml:mi><mml:mn>3</mml:mn></mml:msub><mml:mo>=</mml:mo><mml:mn>0.35</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and computed using the Fokker-Planck equilibria based FPC-LBM. It is seen that the FPC-LBM remains stable even for the cases with an extreme increase in bulk viscosity shown previously, where the MCM-LBM did not remain stable (see <xref ref-type="fig" rid="fluids-09-00255-f010">Figure 10</xref>). This indicates that the FPC-LBM is numerically more stable when compared to the MCM-LBM in such cases as well.</p>
Full article ">Figure 12
<p>Streamlines for three-dimensional lid-driven cubic cavity flow at Reynolds numbers of <inline-formula><mml:math id="mm926"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>100</mml:mn><mml:mo>,</mml:mo><mml:mn>400</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> and 1000 computed using the FPC-LBM along the <inline-formula><mml:math id="mm927"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>a</bold>,<bold>d</bold>,<bold>g</bold>), the <inline-formula><mml:math id="mm928"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>b</bold>, <bold>e</bold>,<bold>h</bold>), and the <inline-formula><mml:math id="mm929"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> centerplane shown in (<bold>c</bold>,<bold>f</bold>,<bold>i</bold>).</p>
Full article ">Figure 13
<p>Comparisons of the horizontal velocity <inline-formula><mml:math id="mm930"><mml:semantics><mml:mrow><mml:mi>u</mml:mi><mml:mo>/</mml:mo><mml:mi>U</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> along the vertical coordinate <inline-formula><mml:math id="mm931"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>/</mml:mo><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm932"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm933"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> (left) and vertical velocity <inline-formula><mml:math id="mm934"><mml:semantics><mml:mrow><mml:mi>v</mml:mi><mml:mo>/</mml:mo><mml:mi>U</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> along the horizontal coordinate <inline-formula><mml:math id="mm935"><mml:semantics><mml:mrow><mml:mi>x</mml:mi><mml:mo>/</mml:mo><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> at <inline-formula><mml:math id="mm936"><mml:semantics><mml:mrow><mml:mi>y</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm937"><mml:semantics><mml:mrow><mml:mi>z</mml:mi><mml:mo>=</mml:mo><mml:mn>0.5</mml:mn><mml:mi>L</mml:mi></mml:mrow></mml:semantics></mml:math></inline-formula> (right) in a three-dimensional cubic cavity flow at different Reynolds numbers computed using the FPC-LBM with the reference results of Ku et al. (1987) [<xref ref-type="bibr" rid="B71-fluids-09-00255">71</xref>] and Shu et al. (2003) [<xref ref-type="bibr" rid="B72-fluids-09-00255">72</xref>]. (<bold>a</bold>,<bold>b</bold>) <inline-formula><mml:math id="mm938"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>100</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>c</bold>,<bold>d</bold>) <inline-formula><mml:math id="mm939"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>400</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>, (<bold>e</bold>,<bold>f</bold>) <inline-formula><mml:math id="mm940"><mml:semantics><mml:mrow><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>=</mml:mo><mml:mn>1000</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula>.</p>
Full article ">Figure 14
<p>Parallel performance of the MPI implementation of our 3D FPC-LBM for lid-driven cubic cavity flow simulations using a grid resolution of <inline-formula><mml:math id="mm941"><mml:semantics><mml:mrow><mml:mn>150</mml:mn><mml:mo>×</mml:mo><mml:mn>150</mml:mn><mml:mo>×</mml:mo><mml:mn>150</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> in our in-house computer cluster.</p>
Full article ">Figure 15
<p>The maximum possible Reynolds number for which the three-dimensional lid-driven cubic cavity flow simulations remain stable at different grid resolutions of <inline-formula><mml:math id="mm942"><mml:semantics><mml:msup><mml:mn>48</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm943"><mml:semantics><mml:msup><mml:mn>64</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula>, <inline-formula><mml:math id="mm944"><mml:semantics><mml:msup><mml:mn>80</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> and <inline-formula><mml:math id="mm945"><mml:semantics><mml:msup><mml:mn>96</mml:mn><mml:mn>3</mml:mn></mml:msup></mml:semantics></mml:math></inline-formula> for different types of LB collision models—raw moments-based MRT-LBM, Maxwellian central moments-based MCM-LBM, factorized central moments-based Factorized LBM, cumulant LBM, and Fokker-Planck central moments-based FPC-LBM.</p>
Full article ">Figure 16
<p>A comparison of the decay rates produced by different LB collision models—SRT-LBM, Maxwellian central moments-based MCM-LBM, cumulant LBM, and Fokker-Planck central moments-based FPC-LBM as compared to the analytically predicted decay rate for the simulation of orthogonal crossing shear waves. Figure (<bold>a</bold>) indicates that the MCM-LBM fails to produce a decay rate similar to that of the analytical solution, and thus is not able to deal with the numerical hyperviscosity effects associated with this problem. Figure (<bold>b</bold>), which is a highly zoomed version of the left figure, indicates that the SRT-LBM can deal with the hyperviscosity effects but also contains unwanted noise. Furthermore, the cumulant LBM and the FPC-LBM are seen to have nearly identical decay rates that are consistent with the analytical solution.</p>
Full article ">Figure 17
<p>Comparisons of turbulence statistics for fully developed turbulent channel flow at a shear Reynolds number of <inline-formula><mml:math id="mm946"><mml:semantics><mml:mrow><mml:msub><mml:mrow><mml:mi>R</mml:mi><mml:mi>e</mml:mi></mml:mrow><mml:mo>*</mml:mo></mml:msub><mml:mo>=</mml:mo><mml:mn>180</mml:mn></mml:mrow></mml:semantics></mml:math></inline-formula> computed using the FPC-LBM and compared to the direct numerical simulations (DNS) data of Lee and Moser (2015) [<xref ref-type="bibr" rid="B74-fluids-09-00255">74</xref>] and experimental data of Kreplin and Eckelmann (1979) [<xref ref-type="bibr" rid="B76-fluids-09-00255">76</xref>]. (<bold>a</bold>) Mean streamwise velocity, (<bold>b</bold>) Root-mean-square (rms) velocity fluctuations, and (<bold>c</bold>) Reynolds stress along the streamwise-wall normal direction.</p>
Full article ">
12 pages, 18868 KiB  
Article
Shape-Memory Effect and the Topology of Minimal Surfaces
by Mengdi Yin and Dimitri D. Vvedensky
Symmetry 2024, 16(9), 1187; https://doi.org/10.3390/sym16091187 - 10 Sep 2024
Viewed by 475
Abstract
Martensitic transformations, viewed as continuous mappings between triply periodic minimal surfaces (TPMSs), as suggested by Hyde and Andersson (Z. Kristallogr. 1986, 174, 225–236), are extended to include paths between the initial and final phases. Reversible transformations, which correspond to [...] Read more.
Martensitic transformations, viewed as continuous mappings between triply periodic minimal surfaces (TPMSs), as suggested by Hyde and Andersson (Z. Kristallogr. 1986, 174, 225–236), are extended to include paths between the initial and final phases. Reversible transformations, which correspond to shape-memory materials, occur only if lattice points remain at flat points on a TPMS throughout a continuous transformation. For the shape-memory material NiTi, the density functional calculations by Hatcher et al. [Phys. Rev. B2009, 80, 144203] yield irreversible and reversible paths with and without energy barriers, respectively, in agreement with our theory. Although there are TPMSs for face-centered and body-centered cubic crystals for iron, the deformation between them is not reversible, in agreement with the non-vanishing energy barriers obtained from the density functional calculations of Zhang et al. (RSC Advances2021, 11, 3043–3048). Full article
(This article belongs to the Section Physics)
Show Figures

Figure 1

Figure 1
<p>Periodic units that repeat along three independent directions, to form the corresponding TPMS: (<b>a</b>) <span class="html-italic">P</span> surface, (<b>b</b>) <span class="html-italic">I</span>-<math display="inline"><semantics> <mrow> <mi>W</mi> <mi>P</mi> </mrow> </semantics></math> surface, (<b>c</b>) <span class="html-italic">F</span>-<math display="inline"><semantics> <mrow> <mi>R</mi> <mi>D</mi> </mrow> </semantics></math> surface, and (<b>d</b>) <span class="html-italic">H</span> surface.</p>
Full article ">Figure 2
<p>Periodic units of (<b>a</b>) <span class="html-italic">P</span> surface deformed to (<b>b</b>) a surface belonging to the <math display="inline"><semantics> <mrow> <mi>o</mi> <mi>P</mi> <mi>a</mi> </mrow> </semantics></math> family, and (<b>c</b>) the top view of the <span class="html-italic">P</span> surface deformed to (<b>d</b>) a surface belonging to the <math display="inline"><semantics> <mrow> <mi>o</mi> <mi>P</mi> <mi>b</mi> </mrow> </semantics></math> family.</p>
Full article ">Figure 3
<p>Atomic arrangements of (<b>a</b>) the Bain transformation path from the FCC to the BCC phase, with an intermediate BCT unit cell (shaded), (<b>c</b>) the NW path, with an intermediate body-centered triclinic unit cell (shaded), (<b>e</b>) the KS transformation path, with an intermediate body-centered triclinic cell, and (<b>g</b>) the BB (OC) path, with an intermediate body-centered triclinic cell depicted in (<b>h</b>). The unit cells are isolated in (<b>b</b>), (<b>d</b>), (<b>f</b>), and (<b>h</b>), respectively. The red and blue spheres indicate atoms in the parent and transformed phases, respectively.</p>
Full article ">Figure 4
<p>The (<b>a</b>) Bain deformation, (<b>b</b>) NW deformation, (<b>c</b>) KS deformation, and (<b>d</b>) BB (OC) deformation. The blue ad spheres signify lattice points in the BCC and FCC cells, respectively. The green and cyan bond lines, which connect the nearest neighbors in the FCC lattice, connect nearest and second-nearest neighbors in the BCC lattice. Therefore, none of these paths is continuous.</p>
Full article ">Figure 5
<p>The NiTi lattice against the <span class="html-italic">P</span> surface. The red and blue spheres are the Ni and Ti atoms, respectively. The magenta spheres are Bravais lattice points. The entire Bravais lattice is shown in (<b>a</b>), with the detail of the positions of the lattice points in (<b>b</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>) B2 NiTi under (<b>b</b>) a single layer <math display="inline"><semantics> <mrow> <mo>〈</mo> <mn>100</mn> <mo>〉</mo> <mo>{</mo> <mn>011</mn> <mo>}</mo> </mrow> </semantics></math> shear and (<b>c</b>) a bilayer <math display="inline"><semantics> <mrow> <mo>〈</mo> <mn>100</mn> <mo>〉</mo> <mo>{</mo> <mn>011</mn> <mo>}</mo> </mrow> </semantics></math> shear. The orange plane in (<b>a</b>) denotes the <math display="inline"><semantics> <mrow> <mo>{</mo> <mn>011</mn> <mo>}</mo> </mrow> </semantics></math> invariant plane; the red and blue spheres denote Ni and Ti atoms, respectively; (<b>b</b>,<b>c</b>) are viewed from the <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>0</mn> <mover> <mn>1</mn> <mo>¯</mo> </mover> <mn>1</mn> <mo>]</mo> </mrow> </semantics></math> direction. In (<b>b</b>,<b>c</b>), <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>,</mo> <mspace width="0.166667em"/> <msub> <mi>l</mi> <mn>2</mn> </msub> </mrow> </semantics></math> are the lengths of cell edges outlined by black dashed lines in the crystal lattice. In (<b>b</b>), <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>1.40</mn> </mrow> </semantics></math>, and in (<b>c</b>) <math display="inline"><semantics> <mrow> <msub> <mi>l</mi> <mn>1</mn> </msub> <mo>/</mo> <msub> <mi>l</mi> <mn>2</mn> </msub> <mo>=</mo> <mn>0.70</mn> </mrow> </semantics></math>. In (<b>b</b>–<b>f</b>), <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <msup> <mn>98</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>. In (<b>d</b>–<b>f</b>), <math display="inline"><semantics> <mrow> <msubsup> <mi>l</mi> <mn>1</mn> <mo>′</mo> </msubsup> <mo>,</mo> <mspace width="0.166667em"/> <msubsup> <mi>l</mi> <mn>2</mn> <mo>′</mo> </msubsup> </mrow> </semantics></math> are the lengths of the cell edges outlined by black dashed lines on the surface. In (<b>d</b>), <math display="inline"><semantics> <mrow> <msubsup> <mi>l</mi> <mn>1</mn> <mo>′</mo> </msubsup> <mo>/</mo> <msubsup> <mi>l</mi> <mn>2</mn> <mo>′</mo> </msubsup> <mo>=</mo> <mn>1.40</mn> </mrow> </semantics></math>; in (<b>e</b>), <math display="inline"><semantics> <mrow> <msubsup> <mi>l</mi> <mn>1</mn> <mo>′</mo> </msubsup> <mo>/</mo> <msubsup> <mi>l</mi> <mn>2</mn> <mo>′</mo> </msubsup> <mo>=</mo> <mn>1.40</mn> </mrow> </semantics></math>; and in (<b>f</b>), <math display="inline"><semantics> <mrow> <msubsup> <mi>l</mi> <mn>1</mn> <mo>′</mo> </msubsup> <mo>/</mo> <msubsup> <mi>l</mi> <mn>2</mn> <mo>′</mo> </msubsup> <mo>=</mo> <mn>0.62</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>(<b>a</b>) Crystal lattice of the 109°–B19′ phase; (<b>b</b>) crystal lattice of the B19′ phase; (<b>c</b>) the TPMS on which the Bravais lattice of the 109°–B19′ phase resides; and (<b>d</b>) the TPMS on which the Bravais lattice of the B19′ lattice resides.</p>
Full article ">Figure 8
<p>(<b>a</b>) Crystal lattice of B2 NiTi, where the red and blue spheres are Ni and Ti atoms, respectively; (<b>b</b>) crystal lattice and Bravais lattice on the <span class="html-italic">P</span> surface viewed from above; (<b>c</b>) deformed crystal lattice of B2 NiTi under the <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>111</mn> <mo>]</mo> </mrow> </semantics></math> elongation; (<b>d</b>) Bravais lattice of R NiTi against the <span class="html-italic">H</span> surface viewed from above; (<b>e</b>) crystal lattice and Bravais lattice against the <span class="html-italic">P</span> surface; (<b>f</b>) Bravais lattice of R NiTi against the <span class="html-italic">H</span> surface viewed from above. The arrow points along the <math display="inline"><semantics> <mrow> <mo>[</mo> <mn>111</mn> <mo>]</mo> </mrow> </semantics></math> direction in the B2 phase.</p>
Full article ">Figure 9
<p>(<b>a</b>) Crystal lattice of the B2 NiTi viewed from the <math display="inline"><semantics> <mrow> <msub> <mrow> <mo>[</mo> <mn>0</mn> <mover> <mn>1</mn> <mo>¯</mo> </mover> <mn>1</mn> <mo>]</mo> </mrow> <mrow> <mi mathvariant="normal">B</mi> <mn>2</mn> </mrow> </msub> <mrow> <mo>]</mo> </mrow> </mrow> </semantics></math> direction; (<b>b</b>) crystal lattice of the B33 NiTi obtained under an alternate bilayer <math display="inline"><semantics> <mrow> <mo>〈</mo> <mn>100</mn> <mo>〉</mo> <mo>{</mo> <mn>011</mn> <mo>}</mo> </mrow> </semantics></math> shear; bilayer marked by a red (resp. blue) dotted rectangle moves along direction denoted by the arrow; in (<b>a</b>,<b>c</b>), <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>109</mn> <mo>.</mo> <msup> <mn>47</mn> <mo>∘</mo> </msup> </mrow> </semantics></math> and in (<b>b</b>,<b>d</b>), <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>107</mn> <mo>.</mo> <msup> <mn>41</mn> <mo>∘</mo> </msup> </mrow> </semantics></math>; (<b>c</b>) plots the crystal lattice (red and blue spheres for Ni and Ti atoms) and Bravais lattice (magenta sphere) of the B2 NiTi against the <span class="html-italic">P</span> surface; (<b>d</b>) plots the Bravais lattice of the B33 NiTi against its corresponding TPMS belonging to the <math display="inline"><semantics> <mrow> <mi>o</mi> <mi>P</mi> <mi>a</mi> </mrow> </semantics></math> family.</p>
Full article ">Figure 10
<p>(<b>a</b>) Crystal lattice (the red and blue spheres represent Ni and Ti atoms, respectively) and Bravais lattice (magenta sphere) of the B2 NiTi against the <span class="html-italic">P</span> surface; (<b>b</b>) Bravais lattice of the B19 NiTi.</p>
Full article ">
17 pages, 8534 KiB  
Article
The Effect of In Concentration and Temperature on Dissolution and Precipitation in Sn–Bi Alloys
by Qichao Hao, Xinfu Tan, Qinfen Gu, Stuart D. McDonald and Kazuhiro Nogita
Materials 2024, 17(17), 4372; https://doi.org/10.3390/ma17174372 - 4 Sep 2024
Viewed by 762
Abstract
Sn–Bi-based, low-temperature solder alloys are being developed to offer the electronics manufacturing industry a path to lower temperature processes. A critical challenge is the significant microstructural and lattice parameter changes that these alloys undergo at typical service temperatures, largely due to the variable [...] Read more.
Sn–Bi-based, low-temperature solder alloys are being developed to offer the electronics manufacturing industry a path to lower temperature processes. A critical challenge is the significant microstructural and lattice parameter changes that these alloys undergo at typical service temperatures, largely due to the variable solubility of Bi during the Sn phase. The influence of alloying additions in improving the performance of these alloys is the subject of much research. This study aims to enhance the understanding of how alloying with In influences these properties, which are crucial for improving the alloy’s reliability. Using in situ heating synchrotron powder X-ray diffraction (PXRD), we investigated the Sn–57 wt% Bi–xIn (x = 0, 0.2, 0.5, 1, 3 wt%) alloys during heating and cooling. Our findings reveal that In modifies the microstructure, promoting more homogeneous Bi distribution during thermal cycling. This study not only provides new insights into the dissolution and precipitation behaviour of Bi in Sn–Bi-based alloys, but also demonstrates the potential of In to improve the thermal stability of these alloys. These innovations contribute significantly to advancing the performance and reliability of Sn–Bi-based, low-temperature solder alloys. Full article
(This article belongs to the Special Issue Electronic Packaging Materials and Technology Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic of the in situ synchrotron PXRD for Sn–57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%) alloy samples.</p>
Full article ">Figure 2
<p>SEM and EDS images of the Sn–57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%) alloys: (<b>a</b>) Sn–57Bi, (<b>b</b>) Sn–57Bi–0.2In, (<b>c</b>) Sn–57Bi–0.5In, (<b>d</b>) Sn–57Bi–1In, (<b>e</b>) Sn–57Bi–3In, and (<b>f</b>) EDS mapping of Sn–57Bi–3In.</p>
Full article ">Figure 3
<p>(<b>a</b>) Cooling curve during solidification of Sn–57Bi–xIn (x = 0,0.2, 0.5, 1, and 3 wt%) alloys, (<b>b</b>) definition of recalescence, (<b>c</b>) eutectic temperature, recalescence during the solidification, (<b>d</b>) primary Sn, and (<b>e</b>) eutectic Sn–Bi.</p>
Full article ">Figure 4
<p>Normalised PXRD patterns for Sn–57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%) alloys at room temperature.</p>
Full article ">Figure 5
<p>Normalised Sn (200) peak during heating for: (<b>a</b>) Sn–57Bi, (<b>b</b>) Sn–57Bi–0.2In, (<b>c</b>) Sn–57Bi–0.5In, (<b>d</b>) Sn–57Bi–1In, and (<b>e</b>) Sn–57Bi–3In, and (<b>f</b>) the splitting temperature and merging temperature for the Sn (200) peak.</p>
Full article ">Figure 6
<p>Normalised Sn (200) peak during heating for (<b>a</b>) Sn–57Bi and (<b>b</b>) Sn–57Bi–3In. Curves have been offset with temperature for clarity.</p>
Full article ">Figure 7
<p>(<b>a</b>) Lattice parameter Sn–a vs. temperature for Sn–57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%), (<b>b</b>) Sn–a changed percentage from 30 °C to 130 °C, and (<b>c</b>) mass fraction of Bi in βSn phase for Sn-57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%), Thermo-calc.</p>
Full article ">Figure 8
<p>(<b>a</b>) Normalised Sn (200) peak during cooling for Sn–57Bi–xIn (x = 0, 0.2, 0.5, 1, and 3 wt%). (<b>b</b>) Schematic of the Bi concentration within the Sn phase.</p>
Full article ">Figure 9
<p>(<b>a</b>) Energy of formation for the dissolution of Bi and In, <span class="html-italic">ΔE<sub>f</sub></span>, vs. In in the Sn–Bi–In alloy and (<b>b</b>) energy of formation for the dissolution of Bi, <span class="html-italic">ΔE<sub>f_Bi</sub></span>, vs. In concentrations in the Sn–Bi–In alloy.</p>
Full article ">
27 pages, 11164 KiB  
Article
Design and Development of a Side Spray Device for UAVs to Improve Spray Coverage in Obstacle Neighborhoods
by Fanrui Kong, Baijing Qiu, Xiaoya Dong, Kechuan Yi, Qingqing Wang, Chunxia Jiang, Xinwei Zhang and Xin Huang
Agronomy 2024, 14(9), 2002; https://doi.org/10.3390/agronomy14092002 - 2 Sep 2024
Viewed by 636
Abstract
Electric multirotor plant protection unmanned aerial vehicles (UAVs) are widely used in China for efficient and precise plant protection at low altitude for low volumes. Unstructured farmland in China has various types of obstacles, and UAVs usually use a detour path to avoid [...] Read more.
Electric multirotor plant protection unmanned aerial vehicles (UAVs) are widely used in China for efficient and precise plant protection at low altitude for low volumes. Unstructured farmland in China has various types of obstacles, and UAVs usually use a detour path to avoid obstacles due to flight altitude limitations. However, existing UAV spray systems do not spray when in obstacle neighborhoods during obstacle avoidance, resulting in insufficient droplet coverage and reduced plant protection quality in the area. To improve the droplet coverage in obstacle neighborhoods, this article carries out a study of side spray technology with an electric quadrotor UAV, and proposes the design and development of a side spray device. The relationship between the obstacle avoidance path of the UAV and the spray pattern of the side spray device and their effect on droplet coverage in obstacle neighborhoods was explored. An accurate measurement method of the relative position between the UAV and obstacles was proposed. Spray angle calculations and nozzle selection for the side spray device were carried out in conjunction with the relative position. A rotor wind field simulation model was designed based on the lattice Boltzmann method (LBM), and the spatial layout of the side spray device on the UAV was designed based on the simulation results. To explore suitable spray patterns for the side spray device, comparative experiments of droplet coverage in obstacle neighborhoods were carried out under different environments, spray patterns, and flight parameter combinations. The relationship between the flight parameter combinations and the distribution uniformity of droplets and the effective swath width of the side spray device was explored. The experimental results were analyzed by an analysis of variance (ANOVA) and a relationship model was obtained. The results showed that the side spray device can effectively improve droplet coverage in obstacle neighborhoods compared to a device without side spray using the same flight parameter combinations. The effective swath width in obstacle neighborhoods can be increased by a minimum of 6.35%, maximum of 35.32%, and average of 15.25% using the side spray device. The error between the predicted values of the relational model and the field experiment results was less than 15%. The results verify the effectiveness and rationality of the method proposed in this article. This study can provide technical and theoretical references for improving the plant protection quality of UAVs in obstacle environments. Full article
Show Figures

Figure 1

Figure 1
<p>Diagram of the UAV structure and obstacle avoidance bypass area: (<b>a</b>) Sketch of UAV structure; (<b>b</b>) rotation direction of rotors; (<b>c</b>) obstacle avoidance path and bypass area.</p>
Full article ">Figure 2
<p>Schematic diagram of relationship between spray angle and obstacle avoidance path.</p>
Full article ">Figure 3
<p>Relative position measurement of the UAV and obstacles.</p>
Full article ">Figure 4
<p>Simplified UAV model and computational domain.</p>
Full article ">Figure 5
<p>Resolution discretization of computational domain.</p>
Full article ">Figure 6
<p>Wind speed sampling point layout diagram.</p>
Full article ">Figure 7
<p>Indoor experiment platform of rotor wind speed: (<b>a</b>) Experiment platform; (<b>b</b>) UAV; (<b>c</b>) Kestrel 4000.</p>
Full article ">Figure 8
<p>Different spray patterns of side spray device: (<b>a</b>) Pattern 1; (<b>b</b>) pattern 2; (<b>c</b>) pattern 3; (<b>d</b>) pattern 4.</p>
Full article ">Figure 9
<p>Side spray device experiment: (<b>a</b>) Non-crop simulation environment experiment; (<b>b</b>) field experiment.</p>
Full article ">Figure 10
<p>Relative position measurement results: (<b>a</b>) Relative position; (<b>b</b>) spray angle required for side spray device nozzle.</p>
Full article ">Figure 11
<p>Nozzle of the side spray device: (<b>a</b>) Spray boundary shrinkage; (<b>b</b>) nozzle object and structure.</p>
Full article ">Figure 12
<p>Simulation value of wind speed at different vertical distance from the rotor.</p>
Full article ">Figure 13
<p>Rotor wind field distribution of the UAV: (<b>a</b>) Tip vortex, spiral vortex, and discretization; (<b>b</b>) contraction and diffusion of rotor wind field.</p>
Full article ">Figure 14
<p>Side spray device spatial layout diagram.</p>
Full article ">Figure 15
<p>Separation and counting of droplets: (<b>a</b>) Original WSP; (<b>b</b>) individual droplet; (<b>c</b>) weakly adherent droplets; (<b>d</b>) strongly adherent droplets.</p>
Full article ">Figure 16
<p>Coverage density of droplet at each sampling point for different spray patterns of the side spray device.</p>
Full article ">
15 pages, 4805 KiB  
Article
Composition-Regulated Photocatalytic Activity of ZnIn2S4@CdS Hybrids for Efficient Dye Degradation and H2O2 Evolution
by Nikolaos Karamoschos, Andreas Katsamitros, Labrini Sygellou, Konstantinos S. Andrikopoulos and Dimitrios Tasis
Molecules 2024, 29(16), 3857; https://doi.org/10.3390/molecules29163857 - 14 Aug 2024
Viewed by 1120
Abstract
Heterostructures of visible light-absorbing semiconductors were prepared through the growth of ZnIn2S4 crystallites in the presence of CdS nanostructures. A variety of hybrid compositions was synthesized. Both reference samples and heterostructured materials were characterized in detail, regarding their morphology, crystalline [...] Read more.
Heterostructures of visible light-absorbing semiconductors were prepared through the growth of ZnIn2S4 crystallites in the presence of CdS nanostructures. A variety of hybrid compositions was synthesized. Both reference samples and heterostructured materials were characterized in detail, regarding their morphology, crystalline character, chemical speciation, as well as vibrational properties. The abovementioned physicochemical characterization suggested the absence of doping phenomena, such as the integration of either zinc or indium ions into the CdS lattice. At specific compositions, the growth of the amorphous ZnIn2S4 component was observed through both XRD and Raman analysis. The development of heterojunctions was found to be composition-dependent, as indicated by the simultaneous recording of the Raman profiles of both semiconductors. The optical band gaps of the hybrids range at values between the corresponding band gaps of reference semiconductors. The photocatalytic activity was assessed in both organic dye degradation and hydrogen peroxide evolution. It was observed that the hybrids demonstrating efficient photocatalytic activity in dye degradation were rather poor photocatalysts for hydrogen peroxide evolution. Specifically, the hybrids enriched in the CdS component were shown to act efficiently for hydrogen peroxide evolution, whereas ZnIn2S4-enriched hybrids demonstrated high potential to photodegrade an azo-type organic dye. Furthermore, scavenging experiments suggested the involvement of singlet oxygen in the mechanistic path for dye degradation. Full article
(This article belongs to the Special Issue Functional Nanomaterials in Green Chemistry, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffractograms of (<b>A</b>) neat CdS, (<b>Β</b>) neat ZIS, (<b>C</b>) “ZIS 99 wt%” hybrid, and (<b>D</b>) “ZIS 16 wt%” hybrid.</p>
Full article ">Figure 2
<p>SEM imaging of (<b>A</b>) neat CdS, (<b>B</b>) neat ZIS, (<b>C</b>) ZIS 99 wt%, and (<b>D</b>) ZIS 16 wt%.</p>
Full article ">Figure 3
<p>(<b>A</b>) Zn2p XPS core level peaks of neat ZIS, ZIS 99 wt%, and ZIS 16 wt% samples; (<b>B</b>) ZnL<sub>3</sub>M<sub>45</sub>M<sub>45</sub> XAES peaks of neat ZIS, ZIS 99 wt% and ZIS 16 wt% samples; (<b>C</b>) S2p XPS core level peaks of neat ZIS; (<b>D</b>) Cd3d XPS core level peaks of ZIS 99 wt% and ZIS 16 wt% samples.</p>
Full article ">Figure 4
<p>Raman analysis of (<b>A</b>) neat ZIS, (<b>B</b>,<b>C</b>) “ZIS 99 wt%”, and (<b>D</b>) “ZIS 16 wt%” samples.</p>
Full article ">Figure 5
<p>Estimation of band gap via the Kubelka–Munk function and DRS data (insets) of (<b>A</b>) neat ZIS; (<b>B</b>) neat CdS; (<b>C</b>) “ZIS 99 wt%” hybrid; and (<b>D</b>) “ZIS 16 wt%” hybrid.</p>
Full article ">Figure 6
<p>(<b>A</b>) Dye degradation profiles for neat CdS, neat ZIS, ZIS 84 wt%, and ZIS 99 wt% samples; (<b>B</b>) hydrogen peroxide evolution profiles for neat CdS, neat ZIS, ZIS 1 wt%, ZIS 4 wt% and ZIS 16 wt% samples; (<b>C</b>) consecutive cycles of dye photodegradation experiments of the ZIS 99 wt% sample; (<b>D</b>) scavenging experiments for dye degradation experiments in the presence of ZIS 99 wt% sample.</p>
Full article ">
10 pages, 3804 KiB  
Article
Study on the Migration Patterns of Oxygen Elements during the Refining Process of Ti-48Al Scrap under Electromagnetic Levitation
by Xinchen Pang, Guifang Zhang, Peng Yan, Zhixiang Xiao and Xiaoliang Wang
Materials 2024, 17(15), 3709; https://doi.org/10.3390/ma17153709 - 26 Jul 2024
Viewed by 563
Abstract
This study investigated the migration patterns of oxygen in the deoxidation process of Ti-48Al alloy scrap using electromagnetic levitation (EML) technology. Scanning electron microscopy (SEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS) were employed to analyze the oxygen distribution patterns and migration [...] Read more.
This study investigated the migration patterns of oxygen in the deoxidation process of Ti-48Al alloy scrap using electromagnetic levitation (EML) technology. Scanning electron microscopy (SEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS) were employed to analyze the oxygen distribution patterns and migration path during EML. The refining process resulted in three types of oxygen migration: (1) escape from the lattice and evaporation in the form of AlO, Al2O; (2) formation of metal oxides and remaining in the alloy melt; (3) attachment to the quartz tube wall in the form of metal oxides such as Al2O3 and Cr2O3. The oxygen content of the scrap was dropped with a deoxidation ratio of 62%. It indicated that EML can greatly promote the migration and removal of oxygen elements in Ti-Al alloy scrap. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic depiction of the EML system.</p>
Full article ">Figure 2
<p>Temperature variation in saturation vapor pressure of Ti, Al, and Cr in Ti-Al alloys.</p>
Full article ">Figure 3
<p>Results of deoxidation experiments under different electromagnetic levitation conditions: (<b>a</b>) melting time, (<b>b</b>) temperature, and (<b>c</b>) initial oxygen content.</p>
Full article ">Figure 4
<p>The EDS detection images of the Ti-Al alloy (<b>a</b>) before EML and (<b>b</b>) under conditions of 2173 K for 40 min.</p>
Full article ">Figure 5
<p>XRD images of the sample before and after EML.</p>
Full article ">Figure 6
<p>The XPS survey spectra of the substances under three different conditions.</p>
Full article ">Figure 7
<p>The XPS detection results for Ti, Al, Cr, and O elements.</p>
Full article ">Figure 8
<p>Deoxidation process schematic of Ti-Al alloys under EML.</p>
Full article ">
10 pages, 4271 KiB  
Communication
Nonlinear Dynamics of Silicon-Based Epitaxial Quantum Dot Lasers under Optical Injection
by Ruilin Fang, Guang-Qiong Xia, Yan-Fei Zheng, Qing-Qing Wang and Zheng-Mao Wu
Photonics 2024, 11(8), 684; https://doi.org/10.3390/photonics11080684 - 23 Jul 2024
Viewed by 783
Abstract
For silicon-based epitaxial quantum dot lasers (QDLs), the mismatches of the lattice constants and the thermal expansion coefficients lead to the generation of threaded dislocations (TDs), which act as the non-radiative recombination centers through the Shockley–Read–Hall (SRH) recombination. Based on a three-level model [...] Read more.
For silicon-based epitaxial quantum dot lasers (QDLs), the mismatches of the lattice constants and the thermal expansion coefficients lead to the generation of threaded dislocations (TDs), which act as the non-radiative recombination centers through the Shockley–Read–Hall (SRH) recombination. Based on a three-level model including the SRH recombination, the nonlinear properties of the silicon-based epitaxial QDLs under optical injection have been investigated theoretically. The simulated results show that, through adjusting the injection parameters including injection strength and frequency detuning, the silicon-based epitaxial QDLs can display rich nonlinear dynamical states such as period one (P1), period two (P2), multi-period (MP), chaos (C), and injection locking (IL). Relatively speaking, for a negative frequency detuning, the evolution of the dynamical state with the injection strength is more abundant, and an evolution path P1-P2-MP-C-MP-IL has been observed. Via mapping the dynamical state in the parameter space of injection strength and frequency detuning under different SRH recombination lifetime, the effects of SRH recombination lifetime on the nonlinear dynamical state of silicon-based epitaxial QDLs have been analyzed. Full article
(This article belongs to the Special Issue Advanced Lasers and Their Applications, 2nd Edition )
Show Figures

Figure 1

Figure 1
<p>QD level structure and carrier dynamics of the three-level model.</p>
Full article ">Figure 2
<p>Calculated SRH recombination lifetime as a function of TD density.</p>
Full article ">Figure 3
<p>Dependence of output photon number on bias current under different values of <span class="html-italic">τ<sub>SRH</sub></span>.</p>
Full article ">Figure 4
<p>Time series, power spectra, and phase portraits of the silicon-based epitaxial QDL under optical injection with Δ<span class="html-italic">ω<sub>inj</sub></span> = −6.3 GHz and <span class="html-italic">k<sub>inj</sub></span> = 0.1 (<b>a1</b>–<b>a6</b>), 0.27 (<b>b1</b>–<b>b6</b>), 0.5 (<b>c1</b>–<b>c6</b>), 1 (<b>d1</b>–<b>d6</b>), 1.3 (<b>e1</b>–<b>e6</b>), 1.4 (<b>f1</b>–<b>f6</b>), and 2 (<b>g1</b>–<b>g6</b>), respectively. Red curves are for ignoring SRH recombination, and blue curves are for <span class="html-italic">τ<sub>SRH</sub></span> = 0.5 ns.</p>
Full article ">Figure 5
<p>Bifurcation diagram of power extreme of peak series as a function of <span class="html-italic">k<sub>inj</sub></span> for the silicon-based epitaxial QDL under optical injection with Δ<span class="html-italic">ω<sub>inj</sub></span> = −6.3 GHz and ignoring SRH recombination (<b>a</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 5 ns (<b>b</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 1 ns (<b>c</b>), and <span class="html-italic">τ<sub>SRH</sub></span> = 0.5 ns (<b>d</b>).</p>
Full article ">Figure 6
<p>Bifurcation diagram of the power extreme of peak series as a function of <span class="html-italic">k<sub>inj</sub></span> for the silicon-based epitaxial QDL under optical injection with Δ<span class="html-italic">ω<sub>inj</sub></span> = −8 GHz and ignoring SRH recombination (<b>a</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 5 ns (<b>b</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 1 ns (<b>c</b>), and <span class="html-italic">τ<sub>SRH</sub></span> = 0.5 ns (<b>d</b>).</p>
Full article ">Figure 7
<p>Mappings of the nonlinear dynamical states distribution of the silicon-based epitaxial QDL in the parameter space of injection strength and frequency detuning under ignoring SRH recombination (<b>a</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 5 ns (<b>b</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 1 ns (<b>c</b>), <span class="html-italic">τ<sub>SRH</sub></span> = 0.5 ns (<b>d</b>).</p>
Full article ">
11 pages, 2013 KiB  
Article
In Situ XRD Measurement for High-Pressure Iron in Laser-Driven Off-Hugoniot State
by Liang Sun, Hao Liu, Xiaoxi Duan, Huan Zhang, Zanyang Guan, Weimin Yang, Xiaokang Feng, Youjun Zhang, Yulong Li, Sanwei Li, Dong Yang, Zhebin Wang, Jiamin Yang, Jin Liu, Wenge Yang, Toshimori Sekine and Zongqing Zhao
Minerals 2024, 14(7), 715; https://doi.org/10.3390/min14070715 - 15 Jul 2024
Viewed by 939
Abstract
The investigation of iron under high pressure and temperatures is crucial to understand the Earth’s core structure and composition and the generation of magnetic fields. Here, we present new in situ XRD measurements for iron in an off-Hugoniot state by laser-driven ramp compression [...] Read more.
The investigation of iron under high pressure and temperatures is crucial to understand the Earth’s core structure and composition and the generation of magnetic fields. Here, we present new in situ XRD measurements for iron in an off-Hugoniot state by laser-driven ramp compression at pressure of 200–238 GPa. The lattice parameters for the hexagonal (hcp)-Fe phase and the c/a ratios were obtained to compare them with previous static and dynamical data, which provides the direct confirmation of such parameters via the different compression paths and strain rates. This work indicates that laser ramp compression can be utilized to provide crystal structure information and direct key information on the crystal structure of Fe at the ultrahigh pressure–temperature conditions relevant for planetology. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup, target configuration and laser profiles in laser ramp compression experiments. (<b>a</b>) Dynamical compression with in situ X-ray diffraction measurement was created by nanosecond lasers. Several beams irradiated the Fe foil to generate the monochromatic X-ray flash for XRD measurement. The XRD snapshot was documented by image plates in the steel diagnostic box when a series of laser pulses drove the sample in the center of box to achieve an extremely high-pressure state. (<b>b</b>) Dynamical compression for the diamond–Fe–diamond layer target with a specific thickness was generated by drive laser beams. A line-imaging velocimetry (VISAR) device monitored the diamond window to reconstruct the off-Hugoniot pressure history of iron under dynamical compression. (<b>c</b>) Laser profile in shot 092. The main driven laser for the sample was the ~12 ns pulse shape (purple curve) with a gradually increasing intensity as ramp compression. A 500 ps square pulse with a 1-nanosecond initial pre-pulse (red curve) at the proper relative time was adopted to create X-ray source. (<b>d</b>) The detected Fe X-ray source emission in shot 092.</p>
Full article ">Figure 2
<p>The X-ray diffraction patterns of iron observed in shot 092 (<b>a</b>) and shot 093 (<b>b</b>). The diffraction patterns were projected into the 2<math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>−</mo> <mi mathvariant="normal">∅</mi> </mrow> </semantics></math> plane after subtracting the background. The blue dashed lines show the peak positions of the reference Pt at ambient conditions for calibration. The red dashed lines show the peaks from compressed Fe as indexed (<span class="html-italic">h l k</span>) for hcp-Fe. The peaks marked by stars (☆) are ghosts from IP faults and background noise.</p>
Full article ">Figure 3
<p>The constructed compression state for iron at high pressure in shot 092. (<b>a</b>) The laser pulses indicate that the X-ray diffraction recording time at ~14 ns was near the time for peak compression. (<b>b</b>) The VISAR data recorded the extracted free surfaces of diamond windows with experimental uncertainties (red, black, green curves). The velocity history shows that the first velocity jump at ~1.5 km/s is followed by ramp loading in ~5 ns to peak compression velocity ~7.5 km/s, and the single-crystal diamond becomes opaque under ramp compression. (<b>c</b>) The constructed stress map in the C-Fe-C target as a function of time by characteristic calculation. The horizontal dashed lines show the boundaries of the diamond ablator, iron sample, and diamond window in the target. The vertical dashed lines show the ~0.5 ns time duration in X-ray diffraction measurement. (<b>d</b>) The averaged stress for the iron sample as a function of time. In this shot, the deduced pressure for iron at the XRD time is 238 GPa.</p>
Full article ">Figure 4
<p>The c/a ratio of hcp-iron at high pressure by static and dynamical compression. The ideal c/a ratio for the hcp structure (grey dashed line) and the typical measurement data (black and red squares) with fitting data in Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] (black dotted, blue dashed, red, blue lines) are plotted for comparison. The ratios from the previous dynamical compression (black, grey, blue circles) from Refs. [<a href="#B24-minerals-14-00715" class="html-bibr">24</a>,<a href="#B27-minerals-14-00715" class="html-bibr">27</a>,<a href="#B49-minerals-14-00715" class="html-bibr">49</a>] are illustrated. Our ramp compression data are between those of previous shock experiments and the static compression data.</p>
Full article ">Figure 5
<p>The volume–c/a ratio relationship for iron and iron alloy at high pressure. The c/a ratios in static compression (squares) from Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] and in the previous dynamical compression (grey, blue circles) from Refs. [<a href="#B27-minerals-14-00715" class="html-bibr">27</a>,<a href="#B49-minerals-14-00715" class="html-bibr">49</a>] are illustrated for comparison. Our data obtained in ramp-compressed iron are illustrated in red circles. The initial volume used for V/V<sub>0</sub> calculation in static compression from Ref. [<a href="#B18-minerals-14-00715" class="html-bibr">18</a>] is 22.468 Å<sup>3</sup>.</p>
Full article ">
16 pages, 9884 KiB  
Article
In-Plane Compression Properties of Continuous Carbon-Fiber-Reinforced Composite Hybrid Lattice Structures by Additive Manufacturing
by Lingqi Jin, Jun Shi, Zhixin Chen, Zhiyang Wang, Yangfan Zhi, Lei Yang and Xinyi Xiao
Polymers 2024, 16(13), 1882; https://doi.org/10.3390/polym16131882 - 1 Jul 2024
Viewed by 1238
Abstract
Continuous-fiber-reinforced composite lattice structures (CFRCLSs) have garnered attention due to their lightweight and high-strength characteristics. Over the past two decades, many different topological structures including triangular, square, hexagonal, and circular units were investigated, and the basic mechanical responses of honeycomb structures under various [...] Read more.
Continuous-fiber-reinforced composite lattice structures (CFRCLSs) have garnered attention due to their lightweight and high-strength characteristics. Over the past two decades, many different topological structures including triangular, square, hexagonal, and circular units were investigated, and the basic mechanical responses of honeycomb structures under various load conditions, including tension, compression, buckling, shear, and fatigue were studied. To further improve the performance of the honeycombs, appropriate optimizations were also carried out. However, the mechanical properties of a single lattice often struggle to exceed the upper limit of its structure. This paper investigates the effect of permutation and hybrid mode on the mechanical properties of CFRCLSs by comparing five structures: rhomboid (R-type), octagon orthogonal array (OOA-type), octagon hypotenuse array (OHA-type), octagon nested array (ONA-type), and rhomboid circle (RC-type), with the conventional hexagonal structure (H-type). CFRCLS samples are fabricated using fused filament fabrication (FFF), with carbon-fiber-reinforced polylactic acid (PLA) as the matrix. The in-plane compression properties, energy absorption characteristics, and deformation behaviors of the hybrid structures were studied by experimental tests. The results demonstrate that different permutation and hybrid modes alter the deformation behaviors and mechanical properties of the structures. Taking elastic modulus as an example, the values of H-type, R-type, OOA-type, OHA-type, ONA-type, and RC-type are, respectively, 6.08 MPa, 5.76 MPa, 19.0 MPa, 10.3 MPa, 31.7 MPa, and 73.2 MPa, while the ratio of their masses is 1:1:1.10:1.52:1.66. Furthermore, hybrid lattice structures exhibit significantly improved mechanical properties compared to single lattice structures. Compared to the single structure R-type, the RC-type increases elastic modulus, yield strength, and energy absorption, respectively, by 12.7 times, 5.4 times, and 4.4 times. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Fused impregnation equipment; (<b>b</b>) Continuous fiber additive manufacturing equipment; (<b>c</b>) Processing diagram.</p>
Full article ">Figure 2
<p>Pictures of printed samples: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type, and structure formation diagram of (<b>g</b>) ONA-type; (<b>h</b>) RC-type.</p>
Full article ">Figure 3
<p>The schematic diagram that an undirected graph transforms into an Eulerian circuit.</p>
Full article ">Figure 4
<p>The designed paths of lattice structures: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type.</p>
Full article ">Figure 5
<p>Experimental equipment (<b>a</b>) and installation of sample (<b>b</b>).</p>
Full article ">Figure 6
<p>The fracture collapse process of lattice structures: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type.</p>
Full article ">Figure 7
<p>The stress–strain curves of lattice structures: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type.</p>
Full article ">Figure 8
<p>The obtained mechanical properties of structures, including (<b>a</b>) Elastic modulus, (<b>b</b>) Yield strength, (<b>c</b>) Elastic limit, (<b>d</b>) Second peak’s stress.</p>
Full article ">Figure 9
<p>Energy absorption per volume of lattice structures: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type.</p>
Full article ">Figure 10
<p>Energy absorption efficiency of lattice structures: (<b>a</b>) H-type; (<b>b</b>) R-type; (<b>c</b>) OOA-type; (<b>d</b>) OHA-type; (<b>e</b>) ONA-type; (<b>f</b>) RC-type.</p>
Full article ">Figure 11
<p>The energy absorption before the densification point (<b>a</b>) and the energy absorption when <math display="inline"><semantics> <mrow> <mi>ε</mi> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math> (<b>b</b>).</p>
Full article ">
48 pages, 1552 KiB  
Article
Three-Dimensional Singular Stress Fields and Interfacial Crack Path Instability in Bicrystalline Superlattices of Orthorhombic/Tetragonal Symmetries
by Reaz A. Chaudhuri
Crystals 2024, 14(6), 523; https://doi.org/10.3390/cryst14060523 - 30 May 2024
Viewed by 453
Abstract
First, a recently developed eigenfunction expansion technique, based in part on the separation of the thickness variable and partly utilizing a modified Frobenius-type series expansion technique in conjunction with the Eshelby–Stroh formalism, is employed to derive three-dimensional singular stress fields in the vicinity [...] Read more.
First, a recently developed eigenfunction expansion technique, based in part on the separation of the thickness variable and partly utilizing a modified Frobenius-type series expansion technique in conjunction with the Eshelby–Stroh formalism, is employed to derive three-dimensional singular stress fields in the vicinity of the front of an interfacial crack weakening an infinite bicrystalline superlattice plate, made of orthorhombic (cubic, hexagonal, and tetragonal serving as special cases) phases of finite thickness and subjected to the far-field extension/bending, in-plane shear/twisting, and anti-plane shear loadings, distributed through the thickness. Crack-face boundary and interface contact conditions as well as those that are prescribed on the top and bottom surfaces of the bicrystalline superlattice plate are exactly satisfied. It also extends a recently developed concept of the lattice crack deflection (LCD) barrier to a superlattice, christened superlattice crack deflection (SCD) energy barrier for studying interfacial crack path instability, which can explain crack deflection from a difficult interface to an easier neighboring cleavage system. Additionally, the relationships of the nature (easy/easy, easy/difficult, or difficult/difficult) interfacial cleavage systems based on the present solutions with the structural chemistry aspects of the component phases (such as orthorhombic, tetragonal, hexagonal, as well as FCC (face-centered cubic) transition metals and perovskites) of the superlattice are also investigated. Finally, results pertaining to the through-thickness variations in mode I/II/III stress intensity factors and energy release rates for symmetric hyperbolic sine-distributed loads and their skew-symmetric counterparts that also satisfy the boundary conditions on the top and bottom surfaces of the bicrystalline superlattice plate under investigation also form an important part of the present investigation. Full article
(This article belongs to the Section Crystal Engineering)
Show Figures

Figure 1

Figure 1
<p>Schematic of a bicrystalline superlattice plate with an interfacial crack.</p>
Full article ">Figure 2
<p>Variation in stress intensity factors through thickness for sine hyperbolic load: (<bold>a</bold>) symmetric and (<bold>b</bold>) skew-symmetric.</p>
Full article ">Figure 3
<p>Variation in (modes I, II, or III) energy release rate through thickness due to far-field sine hyperbolic load.</p>
Full article ">
21 pages, 23893 KiB  
Article
Algorithm Based on Morphological Operators for Shortness Path Planning
by Jorge L. Perez-Ramos, Selene Ramirez-Rosales, Daniel Canton-Enriquez, Luis A. Diaz-Jimenez, Gabriela Xicotencatl-Ramirez, Ana M. Herrera-Navarro and Hugo Jimenez-Hernandez
Algorithms 2024, 17(5), 184; https://doi.org/10.3390/a17050184 - 29 Apr 2024
Cited by 1 | Viewed by 1063
Abstract
The problem of finding the best path trajectory in a graph is highly complex due to its combinatorial nature, making it difficult to solve. Standard search algorithms focus on selecting the best path trajectory by introducing constraints to estimate a suitable solution, but [...] Read more.
The problem of finding the best path trajectory in a graph is highly complex due to its combinatorial nature, making it difficult to solve. Standard search algorithms focus on selecting the best path trajectory by introducing constraints to estimate a suitable solution, but this approach may overlook potentially better alternatives. Despite the number of restrictions and variables in path planning, no solution minimizes the computational resources used to reach the goal. To address this issue, a framework is proposed to compute the best trajectory in a graph by introducing the mathematical morphology concept. The framework builds a lattice over the graph space using mathematical morphology operators. The searching algorithm creates a metric space by applying the morphological covering operator to the graph and weighing the cost of traveling across the lattice. Ultimately, the cumulative traveling criterion creates the optimal path trajectory by selecting the minima/maxima cost. A test is introduced to validate the framework’s functionality, and a sample application is presented to validate its usefulness. The application uses the structure of the avenues as a graph. It proposes a computable approach to find the most suitable paths from a given start and destination reference. The results confirm that this is a generalized graph search framework based on morphological operators that can be compared to the Dijkstra approach. Full article
(This article belongs to the Section Algorithms for Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>Morphological basic operators: (<b>a</b>) dilation operator and (<b>b</b>) erosion operator.</p>
Full article ">Figure 2
<p>Covering dilation operator and the frequency with which each node is dilated.</p>
Full article ">Figure 3
<p>Example of searching algorithm with (<b>a</b>) avenue scenario with node origin and destination; (<b>b</b>) frequency with which dilation is applied; and (<b>c</b>) dilations required to reach the goal.</p>
Full article ">Figure 4
<p>Multiple solutions for best path trajectory. (<b>a</b>) shows the different solutions located; (<b>b</b>) the frequency is illustrated as level curves.</p>
Full article ">Figure 5
<p>Map encoding process: (<b>a</b>) geographical binary map layer; (<b>b</b>) built graph connection; (<b>c</b>) computation of distance transformation; and (<b>d</b>) detection of maximum zones.</p>
Full article ">Figure 6
<p>Example of the searching algorithm in a graph space with (<b>a</b>) an avenue scenario, including the origin and destination nodes, and (<b>b</b>) the frequency with which dilation is applied and the computation of the best path trajectory.</p>
Full article ">Figure 7
<p>Complexity for different <math display="inline"><semantics> <mi>λ</mi> </semantics></math> values, matched with the reference algorithm.</p>
Full article ">
18 pages, 374 KiB  
Article
Combinatorial Generation Algorithms for Directed Lattice Paths
by Yuriy Shablya, Arsen Merinov and Dmitry Kruchinin
Mathematics 2024, 12(8), 1207; https://doi.org/10.3390/math12081207 - 17 Apr 2024
Viewed by 1035
Abstract
Graphs are a powerful tool for solving various mathematical problems. One such task is the representation of discrete structures. Combinatorial generation methods make it possible to obtain algorithms that can create discrete structures with specified properties. This article is devoted to issues related [...] Read more.
Graphs are a powerful tool for solving various mathematical problems. One such task is the representation of discrete structures. Combinatorial generation methods make it possible to obtain algorithms that can create discrete structures with specified properties. This article is devoted to issues related to the construction of such combinatorial generation algorithms for a wide class of directed lattice paths. The main method used is based on the representation of a given combinatorial set in the form of an AND/OR tree structure. To apply this method, it is necessary to have an expression for the cardinality function of a combinatorial set that satisfies certain requirements. As the main result, we have found recurrence relations for enumerating simple directed lattice paths that satisfy the requirements of the applied method and have constructed the corresponding AND/OR tree structure. Applying the constructed AND/OR tree structure, we have developed new algorithms for ranking and unranking simple directed lattice paths. Additionally, the obtained results were generalized to the case of directed lattice paths. Full article
(This article belongs to the Special Issue Advances in Graph Theory: Algorithms and Applications)
Show Figures

Figure 1

Figure 1
<p>All options for reaching the point <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>n</mi> <mo>,</mo> <mi>m</mi> <mo>)</mo> </mrow> </semantics></math> using one step from <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>=</mo> <mo>{</mo> <msub> <mi mathvariant="bold">s</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi mathvariant="bold">s</mi> <mn>2</mn> </msub> <mo>,</mo> <mo>…</mo> <mo>,</mo> <msub> <mi mathvariant="bold">s</mi> <mi>k</mi> </msub> <mo>}</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>All possible (black line) and impossible (red line) steps for a simple directed lattice path that begins at <math display="inline"><semantics> <mrow> <mo>(</mo> <mn>0</mn> <mo>,</mo> <mn>0</mn> <mo>)</mo> </mrow> </semantics></math> and ends at <math display="inline"><semantics> <mrow> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mi>m</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>An example of applying (<a href="#FD1-mathematics-12-01207" class="html-disp-formula">1</a>): (<b>a</b>) without (<a href="#FD2-mathematics-12-01207" class="html-disp-formula">2</a>) and (<a href="#FD3-mathematics-12-01207" class="html-disp-formula">3</a>). (<b>b</b>) with (<a href="#FD2-mathematics-12-01207" class="html-disp-formula">2</a>) and (<a href="#FD3-mathematics-12-01207" class="html-disp-formula">3</a>).</p>
Full article ">Figure 4
<p>An AND/OR tree structure for <math display="inline"><semantics> <msubsup> <mi>W</mi> <mi>n</mi> <mi>m</mi> </msubsup> </semantics></math> based on (<a href="#FD1-mathematics-12-01207" class="html-disp-formula">1</a>).</p>
Full article ">Figure 5
<p>An AND/OR tree structure for <math display="inline"><semantics> <msubsup> <mi>W</mi> <mn>3</mn> <mn>1</mn> </msubsup> </semantics></math> with <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>=</mo> <mo>{</mo> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mo>−</mo> <mn>1</mn> <mo>)</mo> <mo>,</mo> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>0</mn> <mo>)</mo> <mo>,</mo> <mo>(</mo> <mn>1</mn> <mo>,</mo> <mn>1</mn> <mo>)</mo> <mo>}</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>Transformation of north–east lattice paths into simple directed lattice paths.</p>
Full article ">Figure 7
<p>Transformation of Dyck paths into simple directed lattice paths.</p>
Full article ">Figure 8
<p>Transformation of Delannoy paths into directed lattice paths.</p>
Full article ">Figure 9
<p>Transformation of Schroder paths into directed lattice paths.</p>
Full article ">Figure 10
<p>Transformation of Motzkin paths into directed lattice paths.</p>
Full article ">
Back to TopTop