[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (564)

Search Parameters:
Keywords = metastable phase

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 11311 KiB  
Article
Effect of Cooling Rate on α Variant Selection and Microstructure Evolution in TB17 Titanium Alloy
by Guoqiang Shang, Xueping Gan, Xinnan Wang, Jinyang Ge, Chao Li, Zhishou Zhu, Xiaoyong Zhang and Kechao Zhou
Materials 2024, 17(20), 5010; https://doi.org/10.3390/ma17205010 - 13 Oct 2024
Viewed by 585
Abstract
The α variant selection and microstructure evolution in a new metastable β titanium alloy TB17 were studied in depth by DTA, microhardness, XRD, SEM, and EBSD characterization methods. Under the rapid cooling rate conditions (150 °C/min–400 °C/min), only a very small amount of [...] Read more.
The α variant selection and microstructure evolution in a new metastable β titanium alloy TB17 were studied in depth by DTA, microhardness, XRD, SEM, and EBSD characterization methods. Under the rapid cooling rate conditions (150 °C/min–400 °C/min), only a very small amount of granular αWM (α Widmanstatten precipitates within the grains) precipitated within the grains. The secondary α phase precipitated in the alloy changed from granular to fine needle-like at moderate cooling rates (15 °C/min–20 °C/min). When continuing to slow down the cooling rates (10 °C/min and 1 °C/min), the αGB (α precipitates along the β grain boundaries), αWGB (α Widmanstatten precipitates that developed from β grain boundaries or αGB) and αWM grew rapidly. Moreover, the continuous cooling transformation (CCT) diagram illustrated the effect of cooling rate on the β/α phase transition. EBSD analysis revealed that the variants selection of α near the original β grain boundary is mainly divided into three categories. (i) The double-BOR (Burgers orientation relationship) αWGB colonies within neighboring β grains grow in different directions but have the same crystallographic orientation. (ii) The double-BOR αWGB colonies within neighboring β grains have different growth directions and different crystallographic orientations. (iii) The double-BOR αWGB colonies within the same grain have the same growth direction, but different crystallographic directions. And these double-BOR αWGB colonies correspond to two variants of the given {0001}α//{110}β. Full article
(This article belongs to the Special Issue Research on Performance Improvement of Advanced Alloys)
Show Figures

Figure 1

Figure 1
<p>As-forged microstructure of TB17 titanium alloy.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>i</b>) are the relative coefficient curves and dilatometric curves of the alloy at cooling rates of 1 °C/min, 10 °C/min, 15 °C/min, 20 °C/min, 50 °C/min, 100 °C/min, 150 °C/min, 200 °C/min, and 400 °C/min, respectively; the horizontal solid lines in the figure are the relative coefficient curves, and the inclined solid lines are the dilatometric curves.</p>
Full article ">Figure 3
<p>CCT diagram of TB17 titanium alloy.</p>
Full article ">Figure 4
<p>X-ray diffraction results of TB17 titanium alloy under different cooling conditions.</p>
Full article ">Figure 5
<p>The back-scattered electron (BSE) images of TB17 titanium alloy at different cooling conditions. (<b>a</b>) 400 °C/min, (<b>b</b>) 50 °C/min, (<b>c</b>) 20 °C/min, (<b>d</b>) 15 °C/min, (<b>e</b>) 10 °C/min, (<b>f</b>) 1 °C/min.</p>
Full article ">Figure 6
<p>Evolution of microstructures under different cooling rates: (<b>a</b>) rapid cooling rate, (<b>b</b>) intermediate cooling rate, and (<b>c</b>) slow cooling rate.</p>
Full article ">Figure 7
<p>The thickness of α<sub>GB</sub>, α<sub>WGB,</sub> and α<sub>WM</sub> under different cooling rates.</p>
Full article ">Figure 8
<p>Microhardness curves of TB17 titanium alloy at different cooling rates.</p>
Full article ">Figure 9
<p>EBSD observation of specimen under the cooling rate of 1 °C/min. (<b>a</b>) IPF orientation distribution map of α phase, (<b>b</b>) IPF orientation distribution map of β phase, (<b>c</b>) corresponding pole <a href="#materials-17-05010-f009" class="html-fig">Figure 9</a> of four β grains.</p>
Full article ">Figure 10
<p>EBSD observation of specimen at the cooling rate of 1 °C/min. (<b>a</b>) Highlighted inverse pole figure map of α<sub>1</sub>; (<b>b</b>) Corresponding pole figures of α<sub>1</sub>, β<sub>2</sub>, and β<sub>3</sub> grains; (<b>c</b>) Schematic illustration of the growth direction of α<sub>1</sub>.</p>
Full article ">Figure 11
<p>EBSD observation of specimens at the cooling rate of 1 °C/min. (<b>a</b>) Highlighted inverse pole figure map of α<sub>2</sub> and α<sub>3</sub>; (<b>b</b>) Corresponding pole figures of α<sub>2</sub>, α<sub>3</sub>, β<sub>2</sub>, and β<sub>3</sub> grains; (<b>c</b>) Schematic illustration of the growth direction of α<sub>2</sub> and α<sub>3</sub>.</p>
Full article ">Figure 12
<p>EBSD observation of specimen at the cooling rate of 1 °C/min. (<b>a</b>) Highlighted inverse pole figure map of α<sub>4</sub> and α<sub>5</sub>; (<b>b</b>) Corresponding pole figures of α<sub>4</sub>, α<sub>5</sub>, β<sub>1</sub>, and β<sub>3</sub> grains; (<b>c</b>) Schematic illustration of the growth direction of α<sub>4</sub> and α<sub>5</sub>.</p>
Full article ">
12 pages, 2706 KiB  
Article
Chromium Immobilization as Cr-Spinel by Regulation of Fe(II) and Fe(III) Concentrations
by Tianci Hua, Yanzhang Li, Bingxu Hou, Yimei Du, Anhuai Lu and Yan Li
Minerals 2024, 14(10), 1024; https://doi.org/10.3390/min14101024 - 13 Oct 2024
Viewed by 422
Abstract
The complex environmental conditions at Cr-contaminated sites, characterized by uneven ion distribution, oxidants competition, and limited solid-phase mobility, lead to inadequate mixing of Fe-based reducing agents with Cr, posing significant challenges to the effectiveness of Cr remediation through Cr-spinel precipitation. This study investigates [...] Read more.
The complex environmental conditions at Cr-contaminated sites, characterized by uneven ion distribution, oxidants competition, and limited solid-phase mobility, lead to inadequate mixing of Fe-based reducing agents with Cr, posing significant challenges to the effectiveness of Cr remediation through Cr-spinel precipitation. This study investigates the distinct roles of Fe(II), Fe(III), and Cr(III) in Cr-spinel crystallization under ambient temperature and pressure. X-ray diffraction, scanning electron microscopy, transmission electron microscopy, X-ray absorption near-edge structure spectroscopy, and Mössbauer spectroscopy were employed to elucidate the phase composition, microstructure, and ion coordination within the precipitates. Our findings indicate that Fe(II) acts as a catalyst in the formation of the spinel phase, occupying octahedral sites within the spinel structure. Under the catalytic influence of Fe(II), Fe(III) transitions into the spinel phase, occupying both the tetrahedral and the remaining octahedral sites. Meanwhile, Cr(III), due to its high octahedral site preference energy, preferentially occupies the octahedral sites. When Fe(II) or Fe(III) is present but does not meet the ideal stoichiometric ratio, a deficiency in Fe(II) leads to low yield and poor crystallinity of Cr-spinel, whereas a deficiency in Fe(III) can completely inhibit its formation. Conversely, when either Fe(II) or Fe(III) is in excess, the formation of Cr-spinel remains feasible. Furthermore, metastable Cr phases can be transformed into stable Cr-spinel by adjusting the Fe(II)/Fe(III)/Cr(III) ratio. These results highlight the broad range of conditions under which Cr-spinel mineralization can occur in environmental settings, enhancing our understanding of the mechanisms driving Cr-spinel formation in Cr-contaminated sites treated with Fe-based reducing agents. This research provides critical insights for optimizing Cr remediation strategies. Full article
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)
Show Figures

Figure 1

Figure 1
<p>Synthesis of Cr-spinel with Fe(II) at non-ideal stoichiometric ratios. (<b>a</b>) Comparison of the reduction processes of Fe(III) and Cr(III) co-precipitate by <span class="html-italic">S.</span> MR-1. (<b>b</b>) XRD patterns of Fe(II) precipitate, co-precipitate of Fe(III) and Cr(III), reduction of Fe(III) and Cr(III) co-precipitate by <span class="html-italic">S.</span> MR-1, and the co-precipitation of Fe(II), Fe(III), and Cr(III) at stoichiometric ratios of 2:1.5:0.5 and 0.5:1.5:0.5, respectively. The symbols ♥ and ♦ represent spinel and goethite, respectively.</p>
Full article ">Figure 2
<p>XRD patterns of Fe(III) precipitate, co-precipitate of Fe(II) and Cr(III), the oxidation of Fe(II) and Cr(III) co-precipitate by O<sub>2</sub>, and the co-precipitation of Fe(II), Fe(III), and Cr(III) at stoichiometric ratios of 1:2:1 and 1:0.5:1, respectively. The symbols ♥ and ♦ represent spinel and goethite, respectively.</p>
Full article ">Figure 3
<p>XRD patterns of Cr(III) precipitate, co-precipitate of Fe(II) and Fe(III), and the co-precipitation of Fe(II), Fe(III), and Cr(III) at stoichiometric ratios of 1:1.5:0.25 and 1:1.5:2, respectively. The symbols ♥ and ♦ represent spinel and goethite, respectively.</p>
Full article ">Figure 4
<p>XRD patterns of precipitates formed from various mixing sequences of ion precipitates. The symbols ♥ and ♦ represent spinel and goethite, respectively.</p>
Full article ">Figure 5
<p>SEM-EDS and TEM patterns of (<b>a</b>) the mixed precipitate of Fe(III) and Cr(III) co-precipitate with Fe(II) precipitate, (<b>b</b>) the mixed precipitate of Fe(II) and Cr(III) co-precipitate with Fe(III) precipitate, and (<b>c</b>) the mixed precipitate of Fe(II) and Fe(III) co-precipitate with Cr(III) precipitate. The red plus sign indicates the detection position of EDS.</p>
Full article ">Figure 6
<p>Cr K-edge XANES of the precipitates.</p>
Full article ">Figure 7
<p>Mössbauer spectrum analysis of the mixed precipitate of Fe(II) and Cr(III) co-precipitate with Fe(III) precipitate.</p>
Full article ">
16 pages, 2072 KiB  
Review
Chiral, Topological, and Knotted Colloids in Liquid Crystals
by Ye Yuan and Ivan I. Smalyukh
Crystals 2024, 14(10), 885; https://doi.org/10.3390/cryst14100885 - 11 Oct 2024
Viewed by 384
Abstract
The geometric shape, symmetry, and topology of colloidal particles often allow for controlling colloidal phase behavior and physical properties of these soft matter systems. In liquid crystalline dispersions, colloidal particles with low symmetry and nontrivial topology of surface confinement are of particular interest, [...] Read more.
The geometric shape, symmetry, and topology of colloidal particles often allow for controlling colloidal phase behavior and physical properties of these soft matter systems. In liquid crystalline dispersions, colloidal particles with low symmetry and nontrivial topology of surface confinement are of particular interest, including surfaces shaped as handlebodies, spirals, knots, multi-component links, and so on. These types of colloidal surfaces induce topologically nontrivial three-dimensional director field configurations and topological defects. Director switching by electric fields, laser tweezing of defects, and local photo-thermal melting of the liquid crystal host medium promote transformations among many stable and metastable particle-induced director configurations that can be revealed by means of direct label-free three-dimensional nonlinear optical imaging. The interplay between topologies of colloidal surfaces, director fields, and defects is found to show a number of unexpected features, such as knotting and linking of line defects, often uniquely arising from the nonpolar nature of the nematic director field. This review article highlights fascinating examples of new physical behavior arising from the interplay of nematic molecular order and both chiral symmetry and topology of colloidal inclusions within the nematic host. Furthermore, the article concludes with a brief discussion of how these findings may lay the groundwork for new types of topology-dictated self-assembly in soft condensed matter leading to novel mesostructured composite materials, as well as for experimental insights into the pure-math aspects of low-dimensional topology. Full article
(This article belongs to the Special Issue Liquid Crystal Research and Novel Applications in the 21st Century)
Show Figures

Figure 1

Figure 1
<p>Colloids in liquid crystals (LCs). (<b>a</b>) Microscopic structure of a nematic LC with rod-like mesogens, i.e., pentylcyanobiphenyl (5CB). The micrograph (right image) shows the texture of a 5CB droplet observed under a microscope with crossed polarizers, polarization direction marked with white double arrows. Inset shows the chemical structure of 5CB molecules and their collective alignment within a small volume. (<b>b</b>) Topological defects in LCs of different winding numbers; green rods represent LC molecules. (<b>c</b>,<b>d</b>) Homeotropic and planar surface anchoring where LC molecules align perpendicular and parallel to the surface of colloidal inclusions. Orange dots and dashes represent surface functioning agents such as polymer grafting that impose the anchoring direction. (<b>e</b>,<b>f</b>) Micrograhs showing microspheres with homeotropic surface anchoring inducing “hedgehog” point defect and “Saturn ring” line defect; white double arrows indicate the crossed polarizers. (<b>g</b>) Micrographs showing microsphere with planar surface anchoring inducing “boojum” surface defects at the polar points of the sphere. (<b>h</b>–<b>j</b>) Corresponding schematics illustrating LC director field configurations around the colloidal spheres. The black dots and line represent the hedgehog defect, the Saturn ring loop, and the surface boojums, respectively. Schematics are not drawn to scale. The far-field director is shown by the double arrow marked with <b>n</b><sub>0</sub>. Adapted from Ref. [<a href="#B21-crystals-14-00885" class="html-bibr">21</a>].</p>
Full article ">Figure 2
<p>Chiral colloids in LCs. (<b>a</b>) Micrograph of a chiral microstructure obtained by 3D microprinting. (<b>b</b>) Director field distortions around such a particle with planar surface anchoring bound to the confining substrate. The lines in the middle layer and the cylinders show that twist deformation is induced in the director field over the particle despite uniform far-field alignment. (<b>c</b>) Interaction forces vs. distance between the surface-bound chiral structure and a free-floating colloidal sphere. The inset is a micrograph of the interacting objects and the interaction trajectory is color-coded with time. (<b>d</b>,<b>e</b>) Director distributions around right-handed microsprings with a planar surface anchoring at energy-minimized positions. The double arrows indicate the far-field director <b>n</b><sub>0</sub>; the color on the particles represents the orientation of the surface director projected to the plane orthogonal to <b>n</b><sub>0</sub>; color scheme is shown as the inset of (<b>d</b>). (<b>f</b>,<b>g</b>) Snapshots of elasticity-mediated interactions between like- (<b>f</b>) and opposite- (<b>g</b>) handed microsprings, exhibiting attraction and repulsion, respectively, over the time of 10–100 s. Scale bars are 5 μm. (<b>h</b>,<b>i</b>) Numerically calculated Landau-de Gennes free energy vs. particle distances between like- (<b>h</b>) and opposite- (<b>i</b>) handed microsprings. When particles are far away from each other, the free energy scales as <span class="html-italic">d</span><sup>−3</sup> like that of dipole–dipole interactions. The distance between particles <span class="html-italic">d</span> is normalized by the particle radius <span class="html-italic">R</span>; the free energy <span class="html-italic">F</span><sub>LdG</sub> is normalized by the thermal energy <span class="html-italic">k</span><sub>B</sub><span class="html-italic">T</span>, where <span class="html-italic">k</span><sub>B</sub> is the Boltzmann constant and <span class="html-italic">T</span> is the room temperature. Adapted from Refs. [<a href="#B34-crystals-14-00885" class="html-bibr">34</a>,<a href="#B35-crystals-14-00885" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>Topological colloids in LCs. (<b>a</b>,<b>b</b>) A square platelet with a central opening suspended in a LC. The image in (<b>a</b>) is taken with fluorescence confocal polarizing microscopy (FCPM); P<sub>FCPM</sub> indicates the polarization direction of the excitation light. The schematic in (<b>b</b>) shows the director distribution around the platelet. (<b>c</b>,<b>d</b>) Colloidal handlebodies of various genera in LCs. Panels in (<b>c</b>) are micrographs obtained by overlapping fluorescence images with orthogonal excitation polarizations as indicated by the green and magenta arrows; insects below are cross-sectional images in the <span class="html-italic">xz</span> plane taken along the yellow dashed lines. The schematics in (<b>d</b>) represent the director (black lines) distortions and topological defects (red and purple lines; purple dots) induced by the handlebodies; the total topological charge is determined by the particle genus <span class="html-italic">m</span><sub>c</sub> = 1 − <span class="html-italic">g</span>. (<b>e</b>–<b>g</b>) A large torus-shaped colloidal particle with homeotropic surface anchoring suspended in a nematic LC, inducing ½ and edge-pinned ¼ defect lines. The ¼ defect lines (blue lines in the schematic (<b>f</b>,<b>g</b>)) traverses along the edge of the particle and may jump between edges connected by ½ defect lines (red lines in (<b>f</b>,<b>g</b>)). The black arrows in the micrograph in (<b>e</b>) indicate the location of bulk ½ defect lines; tilting of the particle is indicated by the dashed line (rotation axis) and red curved arrow (tilt direction). Insets in (<b>e</b>) are obtained under crossed polarizers; polarization marked by white arrows. (<b>h</b>) Fractal colloidal particles with homeotropic surface anchoring in LCs. The first column is taken at elevated temperatures when the surrounding LC is in isotropic phase; the middle two columns are taken under crossed polarizers (white double arrows) and crossed polarizers with a retardation waveplate (yellow line). The last column is a computer-simulated director field distribution around such particles. Adapted from Refs. [<a href="#B31-crystals-14-00885" class="html-bibr">31</a>,<a href="#B37-crystals-14-00885" class="html-bibr">37</a>,<a href="#B39-crystals-14-00885" class="html-bibr">39</a>,<a href="#B40-crystals-14-00885" class="html-bibr">40</a>].</p>
Full article ">Figure 4
<p>Stimuli-responsive topological colloids. (<b>a</b>) Ring-shaped microparticles made with liquid crystal elastomers change shape upon temperature elevation over the nematic–isotropic phase transition point. The schematic in (<b>a</b>) shows the opening and closing of the central hole, effectively changing the topology of the particle. (<b>b</b>) Alignment and polarization-dependent extinction of plasmonic triangular nanoframes dispersed in a nematic LC. The schematic in (<b>a</b>) shows the director distortions caused by the triangular frame. The normal <b><span class="html-italic">ν</span></b> of the plane containing the nanoframe has the freedom to rotate in a cone shape with the far-field director <b>n</b><sub>0</sub> as the symmetry axis; <b>P</b> represents the polarization of the incident light. (<b>c</b>) Contraction of a ring-shaped particle with homeotropic anchoring and the induced disclination loops (indicated by red lines) in a nematic LC host. The far-field director is perpendicular to the viewing plane. Adapted from Refs. [<a href="#B41-crystals-14-00885" class="html-bibr">41</a>,<a href="#B42-crystals-14-00885" class="html-bibr">42</a>,<a href="#B43-crystals-14-00885" class="html-bibr">43</a>].</p>
Full article ">Figure 5
<p>Knotted (<b>a</b>–<b>e</b>) and linked (<b>f</b>–<b>l</b>) colloids in LCs. (<b>a</b>,<b>b</b>) Optical micrographs of a trefoil knot with planar (<b>a</b>) and homeotropic (<b>b</b>) surface anchoring suspended in LCs. Crossed white double arrows indicate the direction of the polarizer and the analyzer and the yellow double arrow indicates the direction of the slow axis of a 530 nm retardation plate. (<b>c</b>,<b>d</b>) Numerically simulated director field distributions on the surface of trefoil (<b>c</b>) and pentafoil (<b>d</b>) knots with planar surface anchoring. The color represents the orientation of the surface director projected to the plane orthogonal to <b>n</b><sub>0</sub>; the color scheme is shown as the inset of (<b>c</b>). The far-field director <b>n</b><sub>0</sub> is perpendicular to the sample plane as marked by the dot in a circle. (<b>e</b>) Schematic of defect lines (represented by green and magenta lines) induced by and entwined with a trefoil knot of homeotropic anchoring obtained from numerical simulation. (<b>f</b>) Polarizing, fluorescence, and simulated micrographs of linked colloidal rings with tangential surface anchoring suspended in LCs. The green and red double arrows represent the excitation polarization for fluorescence imaging. (<b>g</b>) Elastic interaction energy vs. deviation from the equilibrium center-to-center separation Δ<span class="html-italic">d</span> (black symbols) and orientation Δ<span class="html-italic">α</span> (green symbols) as defined in the inset. (<b>h</b>) Numerically simulated director field distributions on the surface of a Hopf link at the position in (<b>f</b>). The color represents the director orientation as defined in the inset. (<b>i</b>, <b>k</b>) Polarizing micrographs of Hopf (<b>i</b>) and Salomon (<b>k</b>) links with homeotropic surface anchoring suspended in LCs. (<b>j</b>,<b>l</b>) corresponding simulated director field distributions and defect field (represented by the red lines) induced by the links. The red arrow in (<b>j</b>) points at the location where the disclination line jumps from one colloidal loop to the other. The double arrows marked <b>n</b><sub>0</sub> represent the far-field director. Adapted from Refs. [<a href="#B46-crystals-14-00885" class="html-bibr">46</a>,<a href="#B49-crystals-14-00885" class="html-bibr">49</a>].</p>
Full article ">
14 pages, 4685 KiB  
Article
Magnetostrictive Behavior of Severe Plastically Deformed Nanocrystalline Fe-Cu Materials
by Alexander Paulischin, Stefan Wurster, Heinz Krenn and Andrea Bachmaier
Metals 2024, 14(10), 1157; https://doi.org/10.3390/met14101157 - 11 Oct 2024
Viewed by 412
Abstract
Reducing the saturation magnetostriction is an effective way to improve the performance of soft magnetic materials and reduce core losses in present and future applications. The magnetostrictive properties of binary Fe-based alloys are investigated for a broad variety of alloying elements. Although several [...] Read more.
Reducing the saturation magnetostriction is an effective way to improve the performance of soft magnetic materials and reduce core losses in present and future applications. The magnetostrictive properties of binary Fe-based alloys are investigated for a broad variety of alloying elements. Although several studies on the influence of Cu-alloying on the magnetic properties of Fe are reported, few studies have focused on the effect on its magnetostrictive behavior. High pressure torsion deformation is a promising fabrication route to produce metastable, single-phase Fe-Cu alloys. In this study, the influence of Cu-content and the chosen deformation parameters on the microstructural and phase evolution in the Fe-Cu system is investigated by scanning electron microscopy and synchrotron X-ray diffraction. Magnetic properties and magnetostrictive behavior are measured as well. While a reduction in the saturation magnetostriction λs is present for all Cu-contents, two trends are noticeable. λs decreases linearly with decreasing Fe-content in Fe-Cu nanocomposites, which is accompanied by an increasing coercivity. In contrast, both the saturation magnetostriction as well as the coercivity strongly decrease in metastable, single-phase Fe-Cu alloys after HPT-deformation. Full article
(This article belongs to the Special Issue Advances in Magnetic Alloys)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A schematic illustration of the sample preparation for the two-step deformation process with the first deformation step conducted on the big HPT-tool. From the obtained big sample, a strip (green) was cut out, from which the new sample (blue disc) for the second HPT-deformation step was fabricated. (<b>b</b>) A schematic illustration of a bisected small HPT disc with the areas, where measurements were conducted, marked as colored surfaces. The color code of the captions refers to the colored surfaces. The inset shows a sample after HPT-deformation. The black bar is in the size of 10 mm. The orthogonal reference system “axial–radial–tangential” refers to (<b>a</b>,<b>b</b>). The radial positions are highlighted.</p>
Full article ">Figure 2
<p>BSE images of (<b>a</b>) Fe after 20 rotations as well as (<b>b</b>) Fe<sub>95</sub>Cu<sub>5</sub> No. 1, (<b>c</b>) Fe<sub>95</sub>Cu<sub>5</sub> No. 2, (<b>d</b>) Fe<sub>85</sub>Cu<sub>15</sub> No. 1 *, (<b>e</b>) Fe<sub>85</sub>Cu<sub>15</sub> No. 2 and (<b>f</b>) Fe<sub>70</sub>Cu<sub>30</sub> after the second deformation step at a radial position of 2 mm. The Cu-content in at.% determined by EDS is given in the blue boxes. The mean hardness values are given in the green boxes. The scalebar in (<b>a</b>) refers to all microstructures in <a href="#metals-14-01157-f002" class="html-fig">Figure 2</a>. The coordinate system in (<b>a</b>) refers to all subsequent BSE images. All obtained hardness values have been previously published in ref. [<a href="#B25-metals-14-01157" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) BSE images of Fe<sub>95</sub>Cu<sub>5</sub> No. 1 after the second deformation step at a radial position of 0 mm, 1 mm, 2 mm and 3 mm. (<b>b</b>) The corresponding diffraction patterns. The positions and intensities of the reference patterns are indicated as colored bars. Measurements were carried out at DESY PETRA III at beamline P07B.</p>
Full article ">Figure 4
<p>(<b>a</b>) BSE images of Fe<sub>85</sub>Cu<sub>15</sub> No. 1 * after the second deformation step at a radial position of 0 mm, 1 mm, 2 mm and 3 mm. (<b>b</b>) The corresponding diffraction patterns. The positions and intensities of the reference patterns are indicated as colored bars. Measurements were carried out at DESY PETRA III at beamline P21.2.</p>
Full article ">Figure 5
<p>(<b>a</b>) Hysteresis loops determined by SQUID magnetometry of the investigated Fe-Cu system after the second HPT-deformation step. The inset gives a detailed view of the intercepts of the hysteresis loops with the abscissa. (<b>b</b>) Evolution of coercivity <span class="html-italic">H<sub>c</sub></span>, remanence <span class="html-italic">M<sub>r</sub></span> and saturation magnetization <span class="html-italic">M<sub>s</sub></span> in dependence on the Cu-concentration.</p>
Full article ">Figure 6
<p>Saturation magnetostriction <span class="html-italic">λ<sub>s</sub></span> and its dependence on the determined Cu-concentration. The blue, red and magenta triangles are reference values of <span class="html-italic">λ<sub>s</sub></span> for polycrystalline Fe [<a href="#B2-metals-14-01157" class="html-bibr">2</a>,<a href="#B24-metals-14-01157" class="html-bibr">24</a>,<a href="#B27-metals-14-01157" class="html-bibr">27</a>]. The dashed lines, which enclose the gray area, are a linear approximation of <span class="html-italic">λ<sub>s</sub></span> in dependence on Cu-content, following a simple rule of mixture. All values of <span class="html-italic">λ<sub>s</sub></span> have been previously published in ref. [<a href="#B25-metals-14-01157" class="html-bibr">25</a>].</p>
Full article ">
25 pages, 13336 KiB  
Article
Synthesis and Characterization of Ti-13Ta-6Sn Foams Produced Using Mechanical Alloying, the Space Holder Method and Plasma-Assisted Sintering
by Francisco Cavilha Neto, Vagner Kretiska Medeiros, Vicente Salinas-Barrera, Edgar Pio, Claudio Aguilar, Bruno Borges Ramos, Aloísio Nelmo Klein, Bruno Henriques and Cristiano Binder
Metals 2024, 14(10), 1145; https://doi.org/10.3390/met14101145 - 8 Oct 2024
Viewed by 736
Abstract
Highly porous titanium foams are great candidates for replacing bone structures with a low elastic modulus owing to their ability to avoid the stress shielding effect. However, the production of highly porous foams (>70 vol.%) with well-distributed, stable, and predictable porous architectures using [...] Read more.
Highly porous titanium foams are great candidates for replacing bone structures with a low elastic modulus owing to their ability to avoid the stress shielding effect. However, the production of highly porous foams (>70 vol.%) with well-distributed, stable, and predictable porous architectures using powder compaction and space holders is challenging. In this study, pure titanium powder and mechanically alloyed Ti-13Ta-6Sn were mixed with 50, 70, and 80 vol.% KCl powders as a space holder, cold-compacted, and sintered in a plasma-assisted sintering reactor to produce highly porous foams. The space holder was completely removed using heat and plasma species collisions prior to sintering. A Ti-13Ta-6Sn alloy powder with α, β, and metastable FCC-γ phases was synthesized. The characteristics of the alloyed powder, mixing step, and the resulting sintered samples were compared to those of CP-Ti. After sintering, the alloy exhibited α and β phases and a reduced elastic modulus. Foams with an elastic modulus in the range of the cortical and trabecular bones were obtained. The results showed the effects of the space holder volume fractions on the volume fraction, size, distribution, interconnectivity, and shape of the pores. The Ti-13Ta-6Sn foams exhibited a uniform open-celled porous architecture, lower elastic modulus, higher yield strength, and higher passivation resistance than CP-Ti. Ti-13Ta-6Sn exhibited a nontoxic effect for the mouse fibroblast cell line. Full article
(This article belongs to the Special Issue Progress in Biomedical Metallic Materials and Surfaces)
Show Figures

Figure 1

Figure 1
<p>SEM images of the starting powders: (<b>a</b>) CP-Ti at 500×, (<b>b</b>) Ti-13Ta-6Sn at 30,000×, and (<b>c</b>) KCl particles at 500×.</p>
Full article ">Figure 2
<p>X-ray diffraction pattern of milled Ti-13Ta-6Sn powder with Rietveld refinement results.</p>
Full article ">Figure 3
<p>Relative density variation as a function of compaction pressure for powders with and without space holders (SH).</p>
Full article ">Figure 4
<p>KCl particles mixed with (<b>a</b>) CP and (<b>b</b>) Ti-13Ta-6Sn powders, and (<b>c</b>) KCl size and distribution analysis by laser diffraction.</p>
Full article ">Figure 5
<p>Transversal cross-sections of sintered dense (<b>a</b>) CP-Ti and (<b>b</b>) Ti-13Ta-6Sn.</p>
Full article ">Figure 6
<p>Plasma-assisted sintered samples of CP-Ti (<b>a</b>) dense, (<b>b</b>) 50% SH, (<b>c</b>) 70% SH, and (<b>d</b>) 80% SH and Ti-13Ta-6Sn (<b>e</b>) dense, (<b>f</b>) 50% SH, (<b>g</b>) 70% SH, and (<b>h</b>) 80% SH. Images were obtained using a digital camera.</p>
Full article ">Figure 7
<p>Foam cross-sections captured by SEM of CP-Ti with (<b>a</b>) 50%, (<b>b</b>) 70%, and (<b>c</b>) 80% space holders and Ti-13Ta-6Sn with (<b>d</b>) 50%, (<b>e</b>) 70%, and (<b>f</b>) 80% space holders.</p>
Full article ">Figure 7 Cont.
<p>Foam cross-sections captured by SEM of CP-Ti with (<b>a</b>) 50%, (<b>b</b>) 70%, and (<b>c</b>) 80% space holders and Ti-13Ta-6Sn with (<b>d</b>) 50%, (<b>e</b>) 70%, and (<b>f</b>) 80% space holders.</p>
Full article ">Figure 8
<p>Pore SEM-EDS analysis of porous (<b>a</b>) CP-Ti and (<b>b</b>) Ti-13Ta-6Sn.</p>
Full article ">Figure 9
<p>Effect of space holder volume fraction on porosity percentage of CP-Ti and Ti-13Ta-6Sn alloy.</p>
Full article ">Figure 10
<p>Reconstructed 3D images of (<b>a</b>) CP-Ti 50%, (<b>b</b>) CP-Ti 70%, (<b>c</b>) CP-Ti 80%, (<b>d</b>) Ti-13Ta-6Sn 50%, (<b>e</b>) Ti-13Ta-6Sn 70%, and (<b>f</b>) Ti-13Ta-6Sn 80% obtained using X-ray microcomputed tomography.</p>
Full article ">Figure 11
<p>Pore volume fraction per slice of (<b>a</b>) CP-Ti and (<b>b</b>) Ti-13Ta-6Sn alloys.</p>
Full article ">Figure 12
<p>Pore size distributions of (<b>a</b>) CP-Ti and (<b>b</b>) Ti-13Ta-6Sn foams produced with 50, 70, and 80% of SH in volume. Dashed lines indicate total average pore size for each added SH volume fraction.</p>
Full article ">Figure 13
<p>Sphericity distributions of (<b>a</b>) CP-Ti and (<b>b</b>) Ti-13Ta-6Sn alloys.</p>
Full article ">Figure 14
<p>Measured engineering stress–strain curves obtained by compressive tests of (<b>a</b>) all samples, (<b>b</b>) all foams, (<b>c</b>) CP-Ti foams, and (<b>d</b>) Ti-13Ta-6Sn foams.</p>
Full article ">Figure 15
<p>Measured dynamic elastic moduli of dense CP-Ti and Ti-13Ta-6Sn with standard deviation.</p>
Full article ">Figure 16
<p>Measured dynamic elastic moduli of porous CP-Ti and Ti-13Ta-6Sn.</p>
Full article ">Figure 17
<p>Typical potentiodynamic polarization curves obtained from the analyzed samples.</p>
Full article ">Figure 18
<p>Qualitative cytotoxicity analysis of dense Ti-13Ta-6Sn measured each 12 h for a total analysis of 48 h following ISO 10993-5. Black squares indicate the reactivity level at each measurement.</p>
Full article ">
13 pages, 4520 KiB  
Article
Effects of Solidification Thermal Variables on the Microstructure and Hardness of the Silicon Aluminum Bronze Alloy CuAl6Si2
by Paulo Henrique Tedardi do Nascimento, Vinicius Torres dos Santos, Ricardo de Luca, Marcio Rodrigues da Silva, Flavia Goncalves Lobo, Rogerio Teram, Mauricio Silva Nascimento, Ronaldo Camara Cozza, Antonio Augusto Couto, Givanildo Alves dos Santos and Anibal de Andrade Mendes Filho
Metals 2024, 14(10), 1134; https://doi.org/10.3390/met14101134 - 5 Oct 2024
Viewed by 532
Abstract
The properties of the final product obtained by solidification directly result from the thermal variables during solidification. This study aims to analyze the influence of thermal solidification variables on the hardness, microstructure, and phases of the CuAl6Si2 alloy. The material [...] Read more.
The properties of the final product obtained by solidification directly result from the thermal variables during solidification. This study aims to analyze the influence of thermal solidification variables on the hardness, microstructure, and phases of the CuAl6Si2 alloy. The material was solidified using unidirectional solidification equipment under non-stationary heat flow conditions, where heat extraction is conducted through a water-cooled graphite base. The thermal solidification variables were extracted using a data acquisition system, and temperature was monitored at six different positions, with cooling rates ranging from 217 to 3 °C/min from the nearest to the farthest position from the heat extraction point. An optical microscope, scanning electron microscope (SEM), and X-ray diffraction (XRD) were used to verify the fusion structure and determine the volumetric fraction of the formed phases. The XRD results showed the presence of β phases, α phases, and possible Fe3Si2 and Fe5Si3 intermetallics with different morphologies and volumetric fractions. Positions with lower cooling rates showed an increased volume fraction of the α phase and possible intermetallics compared to positions with faster cooling. High cooling rates increased the Brinell hardness of the alloy due to the refined and equiaxed β metastable phase, varying from 143 HB to 126 HB for the highest and lowest rates, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Microstructure development of cast Silicon Aluminum Bronze. Adapted from Ref. [<a href="#B26-metals-14-01134" class="html-bibr">26</a>].</p>
Full article ">Figure 2
<p>Schematic representation of the directional solidification apparatus used to measure the thermal variables of the CuAl<sub>6</sub>Si<sub>2</sub> Alloy. Adapted from Ref. [<a href="#B4-metals-14-01134" class="html-bibr">4</a>].</p>
Full article ">Figure 3
<p>Schematic sequence used for sample cut and preparation.</p>
Full article ">Figure 4
<p>Thermal profiles along the length of the CuAl<sub>6</sub>Si<sub>2</sub> alloy during directional solidification.</p>
Full article ">Figure 5
<p>Solidification thermal variables experimentally obtained for CuAl<sub>6</sub>Si<sub>2</sub> during directional solidification. In (<b>a</b>) the time at which the liquidus temperature (t<sub>L</sub>) is observed in each thermocouple, (<b>b</b>) the velocity liquidus isotherm (V<sub>L</sub>) is observed in each thermocouple, (<b>c</b>) the cooling rate (T<sub>R</sub>) is observed in each thermocouple, and (<b>d</b>) the thermal gradient (G<sub>L</sub>) is observed in each thermocouple.</p>
Full article ">Figure 6
<p>Hardness HB vs. position (<b>left</b>). Microhardness HV1 vs. position (<b>right</b>).</p>
Full article ">Figure 7
<p>Optical microstructural analysis and phase identification via SE-SEM.</p>
Full article ">Figure 8
<p>XRD patterns of the Cu<sub>6</sub>Al<sub>2</sub>Si alloy after directional solidification at different positions.</p>
Full article ">
15 pages, 5312 KiB  
Article
On the Polymorphism of Cu2V2O7: Synthesis and Crystal Structure of δ-Cu2V2O7, a New Polymorph
by Ilya V. Kornyakov and Sergey V. Krivovichev
Crystals 2024, 14(10), 857; https://doi.org/10.3390/cryst14100857 - 29 Sep 2024
Viewed by 556
Abstract
Single crystals of the new modification of copper pyrovanadate, δ-Cu2V2O7, were prepared using the chemical vapor transport reaction method. The crystal structure (monoclinic, P21/n, a = 5.0679(3), b = 11.4222(7), c = [...] Read more.
Single crystals of the new modification of copper pyrovanadate, δ-Cu2V2O7, were prepared using the chemical vapor transport reaction method. The crystal structure (monoclinic, P21/n, a = 5.0679(3), b = 11.4222(7), c = 9.4462(6) Å, β = 97.100(6)°, V = 542.61(6) Å3, Z = 4) was solved by direct methods and refined to R1 = 0.029 for 1818 independent observed reflections. The crystal structure contains two Cu sites: the Cu1 site in [4 + 2]-octahedral coordination and the Cu2 site in [4 + 1]-tetragonal pyramidal coordination. There are two V5+ sites, both tetrahedrally coordinated by O atoms. Two adjacent V1O4 and V2O4 tetrahedra share the O4 atom to form a V2O7 dimer. The crystal structure of δ-Cu2V2O7 can be described as based upon layers of V2O7 dimers of tetrahedra parallel to the (001) plane and interlined by chains of the edge-sharing Cu1O6 and Cu2O5 polyhedra running parallel to the a axis and arranged in the layers parallel to the (001) plane. The crystal chemical analysis of the three other known Cu2V2O7 polymorphs indicates that, by analogy with δ-Cu2V2O7, they are based upon layers of V2O7 groups interlinked by layers consisting of chains of CuOn coordination polyhedra (n = 5, 6). The crystal structures of the Cu2V2O7 polymorphs can be classified according to the mutual relations between the Cu-O chains, on the one hand, and the V2O7 groups, on the other hand. The analysis of the literature data and physical density values suggests that, at ambient pressure, α- and β-Cu2V2O7 are the low- and high-temperature polymorphs, respectively, with the phase transition point at 706–710 °C. The β-phase (ziesite) may form metastably under temperatures below 560 °C and, under heating, transform into the stable α-phase (blossite) at 605 °C. The δ- and γ-polymorphs have the highest densities and most probably are the high-pressure phases. The structural complexity relations among the polymorphs correspond to the sequence α = β < γ < δ; i.e., the δ phase described herein possesses the highest complexity, which supports the hypothesis about its stability under high-pressure conditions. Full article
(This article belongs to the Section Inorganic Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>X-ray powder diffraction pattern of the synthesis mixture consisting of all four polymorphs of Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub>, K<sub>0.5</sub>V<sub>2</sub>O<sub>5</sub>, K<sub>4</sub>CuV<sub>5</sub>O<sub>15</sub> and VOCl. The red line shows calculated X-ray diffraction pattern of the whole mixture; the contribution from δ-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub> is shown by the blue line; open black circles indicate experimentally measured pattern; asterisks shows peaks of unidentified phases.</p>
Full article ">Figure 2
<p>Coordination of Cu<sup>2+</sup> cations (<b>a</b>) and configuration of the V<sub>2</sub>O<sub>7</sub> group (<b>b</b>) in the crystal structure of δ-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub>. Legend: Cu: green; V: orange; O: red. The Cu-O bonds with the Cu-O distances in the ranges of &lt;2.2, 2.2–3.0 and 3.0–3.2 Å are shown as dual-band cylinders, single lines and dashed lines, respectively.</p>
Full article ">Figure 3
<p>The crystal structure of δ-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub> in polyhedral representation (<b>a</b>) and the arrangement of the chains of Cu1O<sub>6</sub> and Cu2O<sub>5</sub> polyhedra within the (001) plane (<b>b</b>). Legend as in <a href="#crystals-14-00857-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 4
<p>The crystal structure of blossite, α-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub>: the Cu coordination (<b>a</b>), the layer of Cu-O chains (<b>b</b>), and the projection of the crystal structure along [01<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mn>1</mn> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>] (<b>c</b>). Legend as in <a href="#crystals-14-00857-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 5
<p>The crystal structure of ziesite, β-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub>: the Cu coordination (<b>a</b>), the layer of Cu-O chains (<b>b</b>), and the projection of the crystal structure along [1<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mn>1</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </semantics></math>] (<b>c</b>). Legend as in <a href="#crystals-14-00857-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 6
<p>The crystal structure of γ-Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub>: the Cu coordination polyhedra (<b>a</b>,<b>b</b>), the layers of Cu-O chains (<b>c</b>,<b>d</b>), and the projection of the crystal structure along the <span class="html-italic">a</span> axis (<b>e</b>). Legend as in <a href="#crystals-14-00857-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 7
<p>The crystal structures of the four Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub> polymorphs viewed as consisting of the Cu-Cu and V-V graphs. Legend as in <a href="#crystals-14-00857-f001" class="html-fig">Figure 1</a>. The two adjacent Cu atoms are linked by dual-banded cylinders and single lines if the Cu⋯Cu distances are in the ranges 2.9–3.0 and 3.1–3.2 Å, respectively. The V-V links correspond to the V⋯V distances in the divanadate V<sub>2</sub>O<sub>7</sub> groups.</p>
Full article ">Figure 8
<p>The structural classification scheme for the four Cu<sub>2</sub>V<sub>2</sub>O<sub>7</sub> polymorphs according to the mutual orientations of Cu-O chains and V<sub>2</sub>O<sub>7</sub> groups.</p>
Full article ">
26 pages, 16057 KiB  
Article
Effect of Residual Stresses on the Fatigue Stress Range of a Pre-Deformed Stainless Steel AISI 316L Exposed to Combined Loading
by Darko Jagarinec and Nenad Gubeljak
Metals 2024, 14(9), 1084; https://doi.org/10.3390/met14091084 - 21 Sep 2024
Viewed by 520
Abstract
AISI 316L austenitic stainless steel is utilized in various processing industries, due to its abrasion resistance, corrosion resistance, and excellent properties over a wide temperature range. The physical and mechanical properties of a material change during the manufacturing process and plastic deformation, e.g., [...] Read more.
AISI 316L austenitic stainless steel is utilized in various processing industries, due to its abrasion resistance, corrosion resistance, and excellent properties over a wide temperature range. The physical and mechanical properties of a material change during the manufacturing process and plastic deformation, e.g., bending. During the combined tensile and bending loading of a structural component, the stress state changes due to the residual stresses and the loading range. To characterize the component’s stress state, the billet was bent to induce residual stress, but a phase transformation to martensite also occurred. The bent billet was subjected to combined tensile–bending and fatigue loading. The experimentally measured the load vs. displacement of the bent billet was compared with the numerical simulations. The results showed that during fatigue loading of the bent billet, both the initial stress state at the critical point and the stress state during the dynamic loading itself must be considered. Analysis was demonstrated only for one single critical point on the surface of the bent billet. The residual stresses due to the phase transformation of austenite to martensite affected the range and ratio of stress. The model for the stress–strain behaviour of the material was established by comparing the experimentally and numerically obtained load vs. displacement curves. Based on the description of the stress–strain behaviour of the pre-deformed material, guidelines have been provided for reducing residual tensile stresses in pre-deformed structural components. Full article
(This article belongs to the Special Issue Fatigue, Fracture and Damage of Steels—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Experimentally obtained engineering and true stress–strain curves.</p>
Full article ">Figure 2
<p>Procedure of the bending process (<b>a</b>) and dimensions of the bending billet (<b>b</b>).</p>
Full article ">Figure 3
<p>Apparatus for bending the billet at the start of bending (<b>a</b>) and end of bending (<b>b</b>).</p>
Full article ">Figure 4
<p>Force displacement loop during the 10 mm bending process, experimental and numerical.</p>
Full article ">Figure 5
<p>Apparatus for combined loading of the bent billet with the distribution of forces and displacements.</p>
Full article ">Figure 6
<p>Decompensation of stress based on the displacement of the bent billet.</p>
Full article ">Figure 7
<p>Schematic prediction of the model for determination of the range of stress for loading with total unloading (R<sub>F</sub> = 0) (<b>a</b>) and loading with partial unloading (R<sub>F</sub> &gt; 0) (<b>b</b>).</p>
Full article ">Figure 8
<p>Finite element model for bending and cyclic loading.</p>
Full article ">Figure 9
<p>Mesh strategy for the finite element model.</p>
Full article ">Figure 10
<p>Distribution of stress through the middle cross-section of the bent billet.</p>
Full article ">Figure 11
<p>Equivalent distribution of plastic strain through the middle cross-section of the bent billet.</p>
Full article ">Figure 12
<p>Distribution of von Mises (<b>a</b>) and S11 (<b>b</b>) residual stress on the bent billet.</p>
Full article ">Figure 13
<p>Analysed paths through the middle cross-section of the bent billet.</p>
Full article ">Figure 14
<p>Distribution of residual stress (<b>a</b>) and PEEQ equivalent plastic strain (<b>b</b>) throughout the middle cross-section.</p>
Full article ">Figure 15
<p>Characteristics of applied force vs. steps of cyclic loading.</p>
Full article ">Figure 16
<p>Distribution of stresses along the inner bending line of the specimen.</p>
Full article ">Figure 17
<p>State of stress at the critical point during cyclic loading with the initial residual stress.</p>
Full article ">Figure 18
<p>Loading characteristics of the bent billet.</p>
Full article ">Figure 19
<p>Load–displacement characteristics for each case of loading.</p>
Full article ">Figure 20
<p>Fractured surface of the bent billet after fatigue loading.</p>
Full article ">Figure 21
<p>Characteristics of moment and force.</p>
Full article ">Figure 22
<p>Determination of the range of loading stress via the maximum and minimum stress obtained by FEM, and by combining the experimental displacement with the numerical simulation of the moment.</p>
Full article ">Figure 23
<p>Fatigue S–N results for the pre-deformed bent billet under three different regimes.</p>
Full article ">Figure A1
<p>The PULSTEC μ-X360n device used for measurement of the residual stresses.</p>
Full article ">Figure A2
<p>Results of measuring the residual stresses at the inside point.</p>
Full article ">Figure A3
<p>Results of measuring the residual stresses at the outside point.</p>
Full article ">
12 pages, 7313 KiB  
Article
Morphological Characteristics of W/Cu Composite Nanoparticles with Complex Phase Structure Synthesized via Reactive Radio Frequency (RF) Thermal Plasma
by Chulwoong Han, Song-Yi Kim, Soobin Kim and Ji-Woon Lee
Metals 2024, 14(9), 1070; https://doi.org/10.3390/met14091070 - 18 Sep 2024
Viewed by 399
Abstract
The W/Cu binary system is characterized by its mutual insolubility and excellent wettability, making W/Cu composite materials ideal for managing thermal and electrical properties in electronic components. To optimize material properties, control over the microstructure is crucial, and nanocomposites with uniform dispersion offer [...] Read more.
The W/Cu binary system is characterized by its mutual insolubility and excellent wettability, making W/Cu composite materials ideal for managing thermal and electrical properties in electronic components. To optimize material properties, control over the microstructure is crucial, and nanocomposites with uniform dispersion offer significant advantages. In this study, W/Cu composite nanoparticles were synthesized by feeding a blended feedstock of tungsten trioxide (WO3) micro-powder and cupric oxide (CuO) micro-powder into a reactive radio frequency (RF) argon–hydrogen thermal plasma system. Cu-coated W nanocomposite particles were obtained through the vaporization, reduction, and condensation processes. The resulting nanocomposite particles were composed of body-centered cubic (BCC) α-W, A15 β-W, and face-centered cubic (FCC) Cu phases, with a chemical composition closely matching theoretical calculations. The phase evolution and morphological changes of the synthesized particles were analyzed as a function of heat treatment temperatures up to 1000 °C in a reducing atmosphere. Up to 600 °C, the phase composition and morphology remained stable. At 800 °C, localized diffusion and coalescence of Cu led to the formation of particulate Cu, and a significant phase transformation from metastable β-W to α-W was observed. Additionally, extensive Cu segregation due to long-range diffusion resulted in distinct Cu-rich and Cu-depleted regions. In these regions, notable sintering of W particles and the complete disappearance of β-W occurred. The results showed that the temperature-dependent redistribution of Cu plays a crucial role in the phase transformation of W and the morphology of W/Cu composite particles. Full article
(This article belongs to the Section Metallic Functional Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of a reactive RF thermal plasma system.</p>
Full article ">Figure 2
<p>(<b>a</b>) STEM image showing morphology, elemental maps of (<b>b</b>) W and (<b>c</b>) Cu, and (<b>d</b>) histogram of Cu weight fraction in W/Cu composite nanoparticles synthesized via RF thermal plasma process.</p>
Full article ">Figure 3
<p>(<b>a</b>) X-ray diffraction pattern showing phase evolution and (<b>b</b>) phase fractions of α-W and β-W with respect to the heat treatment temperature in W/Cu composite nanoparticles synthesized via the RF thermal plasma process.</p>
Full article ">Figure 4
<p>(<b>a</b>) STEM image showing morphology, elemental maps of (<b>b</b>) W, (<b>c</b>) Cu, and (<b>d</b>) histogram of Cu weight fraction in W/Cu composite nanoparticles heat-treated at 800 °C.</p>
Full article ">Figure 5
<p>(<b>a</b>) STEM image showing morphology, elemental maps of (<b>b</b>) W, (<b>c</b>) Cu, and (<b>d</b>) histogram of Cu weight fraction in W/Cu composite nanoparticles heat-treated at 1000 °C.</p>
Full article ">Figure 6
<p>(<b>a</b>) Cross-sectional morphology of Cu segregated region in FIBed sample heat-treated at 1000 °C showing severe segregation of Cu particles and W particle growth. EDS results at the position from 1 to 6 in (<b>a</b>) are represented in (<b>b</b>).</p>
Full article ">Figure 7
<p>Elemental maps showing Cu distribution in (<b>a</b>) as-synthesized nanoparticles, (<b>b</b>) heat-treated nanoparticles at 800 °C, (<b>c</b>) heat-treated nanoparticles at 1000 °C, and (<b>d</b>) a histogram of the Cu weight fraction in W/Cu composite nanoparticles with respect to the heat treatment temperature. Red dots represent W and other colors represent Cu in elemental maps.</p>
Full article ">Figure 8
<p>TEM images of (<b>a</b>) as-synthesized composite particles, (<b>b</b>) heat-treated composite particles at 800 °C, (<b>c</b>) heat-treated composite particles at 1000 °C, and F Elemental maps showing Cu distribution in (<b>a</b>) as-synthesized nanoparticles, (<b>b</b>) heat-treated nanoparticles at 800 °C, (<b>c</b>) heat-treated nanoparticles at 1000 °C, and (<b>d</b>) FE-SEM image of heat-treated nanoparticles at 1000 °C represented for comparison of W particle morphology at Cu-segregated region and Cu-depleted region.</p>
Full article ">Figure 9
<p>Schematic illustration of mechanism for synthesis and sintering of Cu-coated W composite nanoparticles.</p>
Full article ">
15 pages, 353 KiB  
Article
Constraints on Metastable Dark Energy Decaying into Dark Matter
by Jônathas S. T. de Souza, Gustavo S. Vicente and Leila L. Graef
Universe 2024, 10(9), 371; https://doi.org/10.3390/universe10090371 - 18 Sep 2024
Cited by 1 | Viewed by 520
Abstract
We revisit the proposal that an energy transfer from dark energy into dark matter can be described in field theory by a first order phase transition. We analyze a metastable dark energy model proposed in the literature, using updated constraints on the decay [...] Read more.
We revisit the proposal that an energy transfer from dark energy into dark matter can be described in field theory by a first order phase transition. We analyze a metastable dark energy model proposed in the literature, using updated constraints on the decay time of a metastable dark energy from recent data. The results of our analysis show no prospects for potentially observable signals that could distinguish this scenario from the ΛCDM. We analyze, for the first time, the process of bubble nucleation in this model, showing that such model would not drive a complete transition to a dark matter dominated phase even in a distant future. Nevertheless, the model is not excluded by the latest data and we confirm that the mass of the dark matter particle that would result from such a process corresponds to the mass of an axion-like particle, which is currently one of the best motivated dark matter candidates. We argue that extensions to this model, possibly with additional couplings, still deserve further attention as it could provide an interesting and viable description for an interacting dark sector scenario based in a single scalar field. Full article
Show Figures

Figure 1

Figure 1
<p>Shape of the potential <math display="inline"><semantics> <mrow> <mi>V</mi> <mo>(</mo> <mi>φ</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 2
<p>Dependence of the mass on the parameter <math display="inline"><semantics> <mi>α</mi> </semantics></math>, with the latter ranging from <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>10</mn> </mrow> </msup> </semantics></math> to <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>.</p>
Full article ">
30 pages, 8586 KiB  
Review
Unraveling the Dynamic Properties of New-Age Energy Materials Chemistry Using Advanced In Situ Transmission Electron Microscopy
by Subramaniyan Ramasundaram, Sampathkumar Jeevanandham, Natarajan Vijay, Sivasubramani Divya, Peter Jerome and Tae Hwan Oh
Molecules 2024, 29(18), 4411; https://doi.org/10.3390/molecules29184411 - 17 Sep 2024
Viewed by 1034
Abstract
The field of energy storage and conversion materials has witnessed transformative advancements owing to the integration of advanced in situ characterization techniques. Among them, numerous real-time characterization techniques, especially in situ transmission electron microscopy (TEM)/scanning TEM (STEM) have tremendously increased the atomic-level understanding [...] Read more.
The field of energy storage and conversion materials has witnessed transformative advancements owing to the integration of advanced in situ characterization techniques. Among them, numerous real-time characterization techniques, especially in situ transmission electron microscopy (TEM)/scanning TEM (STEM) have tremendously increased the atomic-level understanding of the minute transition states in energy materials during electrochemical processes. Advanced forms of in situ/operando TEM and STEM microscopic techniques also provide incredible insights into material phenomena at the finest scale and aid to monitor phase transformations and degradation mechanisms in lithium-ion batteries. Notably, the solid–electrolyte interface (SEI) is one the most significant factors that associated with the performance of rechargeable batteries. The SEI critically controls the electrochemical reactions occur at the electrode–electrolyte interface. Intricate chemical reactions in energy materials interfaces can be effectively monitored using temperature-sensitive in situ STEM techniques, deciphering the reaction mechanisms prevailing in the degradation pathways of energy materials with nano- to micrometer-scale spatial resolution. Further, the advent of cryogenic (Cryo)-TEM has enhanced these studies by preserving the native state of sensitive materials. Cryo-TEM also allows the observation of metastable phases and reaction intermediates that are otherwise challenging to capture. Along with these sophisticated techniques, Focused ion beam (FIB) induction has also been instrumental in preparing site-specific cross-sectional samples, facilitating the high-resolution analysis of interfaces and layers within energy devices. The holistic integration of these advanced characterization techniques provides a comprehensive understanding of the dynamic changes in energy materials. This review highlights the recent progress in employing state-of-the-art characterization techniques such as in situ TEM, STEM, Cryo-TEM, and FIB for detailed investigation into the structural and chemical dynamics of energy storage and conversion materials. Full article
Show Figures

Figure 1

Figure 1
<p>In situ solid-state/electrochemical biasing TEM characterization for energy materials. (<b>A</b>) Single-particle-level characterization of (<b>a</b>) graphene cage-like layer covering connected with the electrical circuit for external load test analysis, and (<b>b</b>) current-voltage measurements of graphene-encapsulated and amorphous-carbon-coated SiMPs [<a href="#B32-molecules-29-04411" class="html-bibr">32</a>]. (<b>B</b>) Experimental set-up of (<b>a</b>) in situ bias arrangements of tungsten tip and Cu wire inside TEM, (<b>b</b>) low-magnification TEM showing the contact, (<b>c</b>) high-magnification TEM showing the contact between tip, wire, and SEI layer, and (<b>d</b>) TEM showing the surface SEI layer assembly of the integrated set-up, and (<b>e</b>) current-voltage plot indicating critical voltage [<a href="#B33-molecules-29-04411" class="html-bibr">33</a>]. (<b>C</b>) Lithiation of electrochemically biased and (<b>a</b>) arc-discharged MWCNT which is glued to an Al rod (working electrode), Li<sub>2</sub>O grown on Li surface acts as a solid electrolyte, bulk Li metal scratched from tungsten rod acts as counter electrode, (<b>b</b>) pristine MWCNT before coming in contact with Li<sub>2</sub>O/Li electrode, (<b>c</b>) lithiated MWCNT showing uniform Li<sub>2</sub>O layer formation on the surface, and (<b>d</b>–<b>f</b>) corresponding EELS mapping of C, Li, and O, respectively, indicating the nanotube lithiation [<a href="#B34-molecules-29-04411" class="html-bibr">34</a>]. (<b>D</b>) Crack formation in the lithiated Si NWs, (<b>a</b>–<b>i</b>) morphological evolution showing anisotropic elongation and crack during the lithiation of solid cells after contacting the Li<sub>2</sub>O/Li electrode. Red arrows in the panel indicate the propagation of reaction front [<a href="#B38-molecules-29-04411" class="html-bibr">38</a>]. (<b>E</b>) In situ electrochemical functional cell for operando TEM characterization of battery materials [<a href="#B41-molecules-29-04411" class="html-bibr">41</a>]. Figures reproduced with permission from [<a href="#B32-molecules-29-04411" class="html-bibr">32</a>,<a href="#B33-molecules-29-04411" class="html-bibr">33</a>,<a href="#B34-molecules-29-04411" class="html-bibr">34</a>,<a href="#B38-molecules-29-04411" class="html-bibr">38</a>,<a href="#B41-molecules-29-04411" class="html-bibr">41</a>].</p>
Full article ">Figure 2
<p>In situ thermal-induced TEM characterization of energy materials. (<b>A</b>) Thermal welding type experiment of nano solder Sn sheet and Cu structure, showing (<b>a</b>) TEM images of their interface at room temperature (Red, and green parts represent Sn, and Cu, respectively) and (<b>d</b>) after heating (enlarged version (<b>b</b>,<b>e</b>) and HRTEM images (<b>c</b>,<b>f</b>) before and after heating, respectively) [<a href="#B50-molecules-29-04411" class="html-bibr">50</a>]. (<b>B</b>) STEM-HAADF images of samples at all four conditions of the perovskite layer (varying the temperature from 50 to 250 °C) showed no visible changes until 150 °C [<a href="#B1-molecules-29-04411" class="html-bibr">1</a>]. (<b>C</b>) Direct transitions of perovskite structures from tetragonal to trigonal crystalline structure, (<b>a</b>–<b>c</b>) degradation process focusing on single MAPbI<sub>3</sub> grains (scale bar = 2 nm), and (<b>d</b>) MAPbI<sub>3</sub> transition from tetragonal phase to PbI<sub>2</sub> with a trigonal configuration [<a href="#B51-molecules-29-04411" class="html-bibr">51</a>]. (<b>D</b>) Cross-sectional view of in situ analysis showing HAADF images of a perovskite solar cell with a device structure of glass/ITO/TiO<sub>2</sub>/CH<sub>3</sub>NH<sub>3</sub>PbI<sub>3</sub>/spiro-OMeTAD/Ag, and stability check: day 1 shows bright images due to FIB beam focus damage, day 2 shows beam-sensitiveness slowly disappeared, days 7–15 show growth of electron-beam damaged area, and days 20–30 show no further changes (later day-30 samples underwent heating-induced degradation at different time intervals as shown by arrows) [<a href="#B52-molecules-29-04411" class="html-bibr">52</a>]. Figures reproduced with permission from [<a href="#B1-molecules-29-04411" class="html-bibr">1</a>,<a href="#B50-molecules-29-04411" class="html-bibr">50</a>,<a href="#B51-molecules-29-04411" class="html-bibr">51</a>,<a href="#B52-molecules-29-04411" class="html-bibr">52</a>].</p>
Full article ">Figure 3
<p>In situ gas-cell type TEM characterization for energy materials. (<b>A</b>) Schematic demonstration of operando ETEM equipped with windowed gas cell. (<b>B</b>) HAADF-STEM characterization depicting the atomic orientation of Ni/TiO<sub>2</sub> (<b>a</b>) catalyst (exposed to H<sub>2</sub>) prepared under in situ conditions at 400 °C (white arrow: Ni NPs of TiO<sub>2</sub> support, solid lines: TiO<sub>x</sub>-covered Ni atomic planes, and dashed lines: unoccupied facets), (<b>b</b>) strain maps showing atomic displacements, (<b>c</b>) catalyst (exposed to CO<sub>2</sub>:H<sub>2</sub> (0.25 bar:0.75 bar) mixture) at 400 °C showing complete re-exposure of Ni and NPs restructuring, and (<b>d</b>) estimated atomic displacements/reorientations [<a href="#B59-molecules-29-04411" class="html-bibr">59</a>]. (<b>C</b>) Concept of loading ZnO nanowires onto SiN<sub>x</sub> observing windows in a gas-cell setup with built-in MEMS chip (SO<sub>2</sub> atm conditions). (<b>D</b>) In situ TEM imaging taken after exposure to SO<sub>2</sub> gaseous conditions shows (<b>a</b>–<b>h</b>) the nanostructure’s morphological evolution, and (<b>i</b>) its corresponding EDS mapping [<a href="#B61-molecules-29-04411" class="html-bibr">61</a>]. (<b>E</b>) The concept of particle formation and growth observed during the reduction of nickel-phyllosilicate-based catalyst precursor was investigated in an in situ gaseous state. (<b>F</b>) In situ TEM images were observed at different time intervals during the reduction of nickel-phyllosilicates under 1 bar pressure (0.1 sccm gas flow rate at 425 °C) with an electron imaging dose of 30 e<sup>−</sup>·A<sup>−2</sup>·s<sup>−1</sup> showing the nucleation and growth of nanoparticles only in the presence of the electron beam [<a href="#B63-molecules-29-04411" class="html-bibr">63</a>]. Figures reproduced with permission from [<a href="#B59-molecules-29-04411" class="html-bibr">59</a>,<a href="#B61-molecules-29-04411" class="html-bibr">61</a>,<a href="#B63-molecules-29-04411" class="html-bibr">63</a>].</p>
Full article ">Figure 4
<p>In situ liquid-cell/electrochemistry integrated TEM characterization for energy materials. (<b>A</b>) Schematic diagram showing a cross-sectional view of TEM holder, (<b>a</b>) conduit-like silicon nitride membranes encapsulating fluid layer and in situ electrochemical workstation, (<b>b</b>) system of three patterned electrodes with top chip, (<b>c</b>,<b>d</b>) electrochemical activity of Pt cyclic voltammetry (CV) with thick and thin liquid (~150 nm) layers, and (<b>e</b>) Temporal evolution (<b>a</b>–<b>h</b>) occurring in LiFePO<sub>4</sub>/FePO<sub>4</sub> cluster during cycling (charge/discharge). Red, and yellow arrows indicate the propagation of delithiation in core–shell, and left to right pathways [<a href="#B75-molecules-29-04411" class="html-bibr">75</a>]. (<b>B</b>) Schematic representation of (<b>a</b>) in situ liquid-cell nanobattery setup analyzing the lithiation process, (<b>b</b>) SEM images of the electrochemically biased chip with (<b>b</b>) inner side, (<b>c</b>) its magnified view, and (<b>d</b>) Si NW electrode welded onto the Pt contact [<a href="#B76-molecules-29-04411" class="html-bibr">76</a>]. (<b>C</b>) Graphene liquid-cell (GLC) TEM illustration: (<b>a</b>) Si NPs were immersed in liquid electrolyte and placed between graphene layers in a sandwich structure, (<b>b</b>) the whole assembly was mounted on a holey amorphous carbon TEM grid (SEM image, scale bar: 1 μm), and (<b>c</b>) STEM mapping images of O, C, Si, P, and F in the GLC (scale bar: 100 nm) [<a href="#B77-molecules-29-04411" class="html-bibr">77</a>]. (<b>D</b>) Schematic diagram of in situ liquid-cell visualization of (<b>a</b>) MOS<sub>2</sub> reaction on Ti electrodes, (<b>b</b>) assembled cell window area for capturing the dynamic lithiation/de-lithiation process, (<b>c</b>) schematic diagram for nanobeam diffraction characterization on the SEI layer or residual MoS<sub>2</sub> reaction after the process, and (<b>d</b>) typical example showing bright-field or dark-field image reconstruction of the diffraction pattern [<a href="#B78-molecules-29-04411" class="html-bibr">78</a>]. Figures reproduced with permission from [<a href="#B75-molecules-29-04411" class="html-bibr">75</a>,<a href="#B76-molecules-29-04411" class="html-bibr">76</a>,<a href="#B77-molecules-29-04411" class="html-bibr">77</a>,<a href="#B78-molecules-29-04411" class="html-bibr">78</a>].</p>
Full article ">Figure 5
<p>In situ light-induced TEM characterization of energy materials. (<b>A</b>) Schematic representation of integrated optical fiber setup with TEM holder [<a href="#B91-molecules-29-04411" class="html-bibr">91</a>]. (<b>B</b>) Customized fiber connection with cut fiber end projecting inside the microscope vacuum with a 15° angle to avoid optical loss, and the opposite angle cut by 30° to produce a beam that can illuminate the TEM sample [<a href="#B92-molecules-29-04411" class="html-bibr">92</a>]. (<b>C</b>) Schematic representation of in situ light-induced TEM with an integrated gas flow and heating controller inside the compact sample chamber. (<b>D</b>) In situ HRTEM imaging of anatase nanocrystals at 150 °C with/without 1 Torr water pressure: (<b>a</b>–<b>f</b>) diverse experimental conditions starting from no water to fresh water presence even after 40 h in a water/gas environment before exposure to the electron beam [<a href="#B93-molecules-29-04411" class="html-bibr">93</a>]. (<b>E</b>) TEM holder modification with liquid-cell chip arrangement. (<b>a</b>,<b>c</b>,<b>d</b>) schematic representations, (<b>b</b>) liquid cell, and (<b>e</b>) real photograph. (<b>F</b>) HRTEM analysis depicting transition in the morphology of Cu<sub>2</sub>O samples observed at different (<b>a</b>–<b>c</b>) irradiation time intervals such as 1 h, 2 h, 3 h, and (<b>d</b>) schematic showing its evolution over time [<a href="#B95-molecules-29-04411" class="html-bibr">95</a>]. (<b>G</b>) Schematic diagram displaying in situ fabrication of TiO<sub>2</sub>/CdSe nanowire QD solar cell integrated with LED and electrical measurement system in the STM-TEM [<a href="#B8-molecules-29-04411" class="html-bibr">8</a>]. (<b>H</b>) Light-induced rapid phase transformation (through <span class="html-italic">α</span> phase and hydrogen-rich <span class="html-italic">β</span> phase) reaction in the antenna–reactor configuration with illumination at 690 nm (scale bars: 50 nm) visualized over different short time intervals: (<b>a</b>,<b>b</b>) phase transformation at difffernt locations., in both cases (I–V) represents snapshots taken at different time (s) interval. Figures reproduced with permission from [<a href="#B8-molecules-29-04411" class="html-bibr">8</a>,<a href="#B91-molecules-29-04411" class="html-bibr">91</a>,<a href="#B92-molecules-29-04411" class="html-bibr">92</a>,<a href="#B93-molecules-29-04411" class="html-bibr">93</a>,<a href="#B95-molecules-29-04411" class="html-bibr">95</a>,<a href="#B97-molecules-29-04411" class="html-bibr">97</a>].</p>
Full article ">Figure 6
<p>In situ Cryo-TEM characterization of MOF nanostructures. (<b>A</b>) Schematic representation of Cryo-TEM technique for MOF nanostructures for monitoring its dynamics under cryogenic conditions [<a href="#B100-molecules-29-04411" class="html-bibr">100</a>]. (<b>B</b>) HRTEM images of (<b>a</b>) MOF-525(Pt) and (<b>b</b>) PCN-224(Pt) (insets: FFT depicting the predominant lattice fringes) [<a href="#B102-molecules-29-04411" class="html-bibr">102</a>]. (<b>C</b>) HRTEM images of (<b>a</b>) intact MOF-5 crystals with inset FFT corresponding to [<a href="#B100-molecules-29-04411" class="html-bibr">100</a>] zone axis orientation as illustrated in (<b>b</b>); the selected inset from the filtered version (<b>c</b>) was simulated noise-free with (<b>d</b>) HRTEM image (thickness 15.5 nm, Δf = +12 nm, Cs +20 μm), and (<b>e</b>) illustrated structure of MOF-5 crystals (white stripes: terephthalate linker, grey color: pores) [<a href="#B103-molecules-29-04411" class="html-bibr">103</a>]. Figures reproduced with permission from [<a href="#B100-molecules-29-04411" class="html-bibr">100</a>,<a href="#B102-molecules-29-04411" class="html-bibr">102</a>,<a href="#B103-molecules-29-04411" class="html-bibr">103</a>].</p>
Full article ">Figure 7
<p>In situ Cryo-TEM characterization for SEI layer and Li metal interfaces in rechargeable batteries. (<b>A</b>) The method for preserving and stabilizing Li metal by Cryo-TEM whereby the specimen was placed onto the Cryo-TEM holder while maintaining cryogenic conditions, and during insertion of TEM column the temperature is maintained not above −170 °C. (<b>B</b>) Cryo-TEM observations (<b>a</b>) depicting kinked Li metal dendrite and SEI that changes from &lt;211&gt; to &lt;110&gt; and finally comes back to &lt;211&gt; growth lattice projection, (<b>b</b>) Li deposition on Cu TEM grid and storage under cryogenic conditions, (<b>c</b>) atomic-resolution images showing the transition from (i) &lt;211&gt; to &lt;110&gt; and (ii) &lt;110&gt; changes back to &lt;211&gt; lattice orientation, (<b>d</b>) mosaic model, and (<b>e</b>) multi-layer model of dendritic Li deposition on SEI under different carbonate electrolyte conditions [<a href="#B101-molecules-29-04411" class="html-bibr">101</a>]. (<b>C</b>) Detailed representation of Li dissolution under (<b>a</b>,<b>c</b>,<b>e</b>) mosaic SEI, and (<b>b</b>,<b>d</b>,<b>f</b>) multi-layer SEIs [<a href="#B3-molecules-29-04411" class="html-bibr">3</a>]. (<b>D</b>) Structural composition and elemental mapping of Li dendrites formed at the interface; the morphology was differentiated and analyzed under (<b>a</b>,<b>b</b>) Cryo-FIB, (<b>c</b>,<b>d</b>) Cryo-STEM, and (<b>e</b>,<b>f</b>) Cryo-EELS for type I dendrites and type II dendrites, respectively. Scale bar in (<b>a</b>,<b>b</b>), and (<b>c</b>–<b>f</b>) are equal to 1 µM, and 300 nm, respectively [<a href="#B4-molecules-29-04411" class="html-bibr">4</a>]. Figures reproduced with permission from [<a href="#B3-molecules-29-04411" class="html-bibr">3</a>,<a href="#B4-molecules-29-04411" class="html-bibr">4</a>,<a href="#B101-molecules-29-04411" class="html-bibr">101</a>].</p>
Full article ">Figure 8
<p>In situ Cryo-TEM characterization of electron-beam sensitive perovskites. (<b>A</b>) Schematic representation displaying (<b>a</b>) integrated galvanostatic and current-voltage measurements under Cryo-TEM conditions, and (<b>b</b>) five galvanostatic curves between mixed and only electronic states measured at the temperature range from 100 to 295 K. (<b>B</b>) Dynamic changes in optical images of perovskite films from (<b>a</b>–<b>c</b>) dark conditions to 20 mW cm<sup>−2</sup> illumination, (<b>d</b>) schematic illustration depicting the migration of cations MA<sup>+</sup> and anions I<sup>−</sup> through the film [<a href="#B105-molecules-29-04411" class="html-bibr">105</a>]. (<b>C</b>) Drop casting of (<b>a</b>) pristine or UV/moisture-exposed perovskite nanowires onto the TEM grid in nitrogen glove box conditions, in comparison with using (<b>b</b>) Cryo-TEM showing preserved structures, (<b>c</b>) conventional TEM techniques showing electron beam-damaged structures with atomic-level resolution in lattice fringes, (<b>d</b>) TEM image of MAPI<sub>3</sub> NWs after exposing to electron beam, and (<b>e</b>) decomposition of MAPI<sub>3</sub> in to PBI<sub>2</sub> [<a href="#B106-molecules-29-04411" class="html-bibr">106</a>]. Figures reproduced with permission from [<a href="#B105-molecules-29-04411" class="html-bibr">105</a>,<a href="#B106-molecules-29-04411" class="html-bibr">106</a>].</p>
Full article ">Scheme 1
<p>Overview of diverse methodologies adapted for in situ TEM characterization of energy materials.</p>
Full article ">
15 pages, 7673 KiB  
Article
Tensile Deformation Mechanism of an In Situ Formed Ti-Based Bulk Metallic Glass Composites
by Haiyun Wang, Na Chen, Huanwu Cheng, Yangwei Wang and Denghui Zhao
Materials 2024, 17(18), 4486; https://doi.org/10.3390/ma17184486 - 12 Sep 2024
Viewed by 385
Abstract
Ti-based bulk metallic glass composites (BMGMCs) containing an in situ formed metastable β phase normally exhibit enhanced plasticity attributed to induced phase transformation or twinning. However, the underlying deformation micromechanism remains controversial. This study investigates a novel deformation mechanism of Ti-based BMGMCs with [...] Read more.
Ti-based bulk metallic glass composites (BMGMCs) containing an in situ formed metastable β phase normally exhibit enhanced plasticity attributed to induced phase transformation or twinning. However, the underlying deformation micromechanism remains controversial. This study investigates a novel deformation mechanism of Ti-based BMGMCs with a composition of Ti42.3Zr28Cu8.3Nb4.7Ni1.7Be15 (at%). The microstructures after tension were analyzed using advanced electron microscopy. The dendrites were homogeneously distributed in the glassy matrix with a volume fraction of 55 ± 2% and a size of 1~5 μm. The BMGMCs deformed in a serrated manner with a fracture strength (σf) of ~1710 MPa and a fracture strain of ~7.1%, accompanying strain hardening. The plastic deformation beyond yielding was achieved by a synergistic action, which includes shear banding, localized amorphization and a localized BCC (β-Ti) to HCP (α-Ti) structural transition. The localized amorphization was caused by high local strain rates during shear band extension from the amorphous matrix to the crystalline reinforcements. The localized structural transition from BCC to HCP resulted from accumulating concentrated stress during deformation. The synergistic action enriches our understanding of the deformation mechanism of Ti-based BMGMCs and also sheds light on material design and performance improvement. Full article
(This article belongs to the Special Issue Synthesis, Sintering, and Characterization of Composites)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Back scattered SEM image showing dendritic reinforcements were uniformly distributed in the glass matrix; (<b>b</b>) XRD pattern of the as-cast Ti-based BMGMCs showing crystalline peaks were overlapped with amorphous hump.</p>
Full article ">Figure 2
<p>TEM and HRTEM images of the as-cast Ti-BMGMCs: (<b>a</b>) bright field TEM image, selected area electron diffraction (SAED) pattern of (<b>b</b>) β-Ti dendrites, (<b>c</b>) glass matrix, HRTEM images of (<b>d</b>) the interface between β-Ti dendritic phase and glassy phase, (<b>e</b>) β-Ti dendrites and (<b>f</b>) glass matrix.</p>
Full article ">Figure 3
<p>(<b>a</b>) Room-temperature compressive stress–strain curve; and (<b>b</b>) room-temperature tensile stress–strain curve and its corresponding work-hardening rate–strain curve of the Ti-based amorphous matrix composites.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD pattern of the Ti-based BMGMCs before and after tension deformation; SEM images of (<b>b</b>) fracture surface of the specimen after tension in low magnification, (<b>c</b>) fracture surface morphology of dimples and (<b>d</b>) fracture surface morphology of river-like pattern in high magnification.</p>
Full article ">Figure 5
<p>SEM images of tensile deformed Ti-based BMGMCs at room temperature: (<b>a</b>,<b>b</b>) lateral side images in low magnification; (<b>c</b>) profuse shear band near fracture; (<b>d</b>) detailed shear banding in high magnification; (<b>e</b>) deformation feature in dual phase region; (<b>f</b>) magnified image of Zone I showing multiple shear bands; (<b>g</b>) magnified image of Zone II showing shear steps; (<b>h</b>) interactions of shear bands and slip bands due to severe plastic deformation.</p>
Full article ">Figure 6
<p>TEM micrographs in low magnification of the tension-deformed Ti-based BMGMCs: (<b>a</b>) deformed microstructures at low magnification; (<b>b</b>) SAED pattern from amorphous phase; (<b>c</b>) SAED pattern from dendritic crystalline phase; (<b>d</b>) slip step formed in the interface between dendrites and glass matrix.</p>
Full article ">Figure 7
<p>(<b>a</b>) HRTEM images showing slip step formed in the interface between the amorphous phase and dendritic phase; (<b>b</b>) enlarged area from the white square in (<b>a</b>,<b>c</b>) FFTs of red rectangle regions in (<b>b</b>), which follow the directions indicated by the arrows; (<b>d</b>) the curves of true stress vs. time and true strain vs. time; (<b>e</b>) the plot of shear strain rate <math display="inline"><semantics> <mover accent="true"> <mi>γ</mi> <mo>˙</mo> </mover> </semantics></math> vs. time.</p>
Full article ">Figure 8
<p>(<b>a</b>) HRTEM image of tension-fractured Ti-based BMGMCs showing microstructural characteristics in the crystalline region close to slip step; (<b>b</b>) enlarged image of the red square in (<b>a</b>), which shows atomic microstructure of the dendrite phase close to slip step; (<b>c</b>) FFTs of Zone I, Zone II and Zone III indicated by red squares in (<b>b</b>), respectively.</p>
Full article ">Figure 9
<p>(<b>a</b>) An HRTEM image of the region showing a BCC to HCP transition and inset shows a corresponding FFT pattern; (<b>b</b>) IFFT of (<b>a</b>) shows a clear transition from BCC to HCP; (<b>c</b>) a schematic diagram of the transition from BCC to HCP.</p>
Full article ">Figure 10
<p>Schematic illustrations of the microstructural evolution for Ti-based BMGMCs during different tension deformation stages: (<b>a</b>,<b>b</b>) elastic stage, during which dendrites deformed initially, leading to interfacial stress concentration; (<b>c</b>–<b>f</b>) plastic stage, both dendrites and glassy matrix deformed, shear bands formed and propagated, leading to the final failure.</p>
Full article ">
13 pages, 5969 KiB  
Article
Abnormal Effect of Al on the Phase Stability and Deformation Mechanism of Ti-Zr-Hf-Al Medium-Entropy Alloys
by Penghao Yuan, Lu Wang, Ying Liu and Xidong Hui
Metals 2024, 14(9), 1035; https://doi.org/10.3390/met14091035 - 11 Sep 2024
Viewed by 492
Abstract
Complex concentrated alloys, including high-entropy alloys (HEAs) and medium-entropy alloys (MEAs), offer another pathway for developing metals with excellent mechanical properties. However, HEAs/MEAs of different structures often suffer from various drawbacks. So, investigations on the effect of phase and microstructure on their properties [...] Read more.
Complex concentrated alloys, including high-entropy alloys (HEAs) and medium-entropy alloys (MEAs), offer another pathway for developing metals with excellent mechanical properties. However, HEAs/MEAs of different structures often suffer from various drawbacks. So, investigations on the effect of phase and microstructure on their properties become necessary. In the present work, we adjust the phase constitution and microstructure by Al addition in a series of (Ti2ZrHf)100−xAlx (x = 12, 14, 16, 18, 20, at.%, named Alx) MEAs. Different from traditional titanium, Al shows a β-stabilizing effect, and the phase follows the evolution of α′(α)→α″→β + ω + B2 with Al increasing from 12 to 20 at.%, which could not be predicted by the CALPHAD (Calculate Phase Diagrams) method or the Bo-Md diagram because of the complex interactions among composition elements. At a low Al content, the solid solution strengthening of the HCP phase contributes to the extremely high strength with a σ0.2 of 1528 MPa and σb of 1937 MPa for Al14. The appearance of α″ deteriorates the deformation capability with increasing Al content in the Al16 and Al18 MEAs. In the Al20 MEA, Al improves the formations of ordered B2 and metastable β. The phase transformation strengthening, including B2 to BCC and BCC to α″, together with the precipitation strengthening of ω, brings about a high work-hardening ratio (above 5 GPa) and improvements in ductility (6.8% elongation). This work provides guidelines for optimizing the properties of MEAs. Full article
(This article belongs to the Section Entropic Alloys and Meta-Metals)
Show Figures

Figure 1

Figure 1
<p>The phase content predicted by Thermo-calc. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20. The legend in (<b>e</b>) applies to (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 2
<p>The phase prediction by the Bo-Md method.</p>
Full article ">Figure 3
<p>The XRD diffractions of Ti-Zr-Hf-Al MEAs.</p>
Full article ">Figure 4
<p>The microstructure of Ti-Zr-Hf-Al MEAs. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20.</p>
Full article ">Figure 5
<p>The microstructure investigated by TEM for the Al12 MEA. (<b>a</b>) The HAADF image and EDS elements mapping; (<b>b</b>) typical bright-field TEM images showing the laths with different sizes in the Al12 alloy, together with the inset image of SAED for the region marked by the red circle; (<b>c</b>) the HRTEM image of the area marked by the yellow square in (<b>c</b>); (<b>d</b>) the FFT of (<b>c</b>); the FFT (<b>e1</b>) and amplified image (<b>e2</b>) for the area e in (<b>c</b>); and the FFT (<b>f1</b>) and amplified image (<b>f2</b>) for the area f in (<b>c</b>).</p>
Full article ">Figure 6
<p>The microstructure investigated by TEM for the Al20 MEA. (<b>a</b>) The HAADF image and EDS element mapping; (<b>b1</b>) typical bright-field TEM image for the Al20 alloy; (<b>b2</b>) SAED for the region marked by the red circle in (<b>b1</b>); (<b>b3</b>) the dark-field TEM image of the ω phase shown in (<b>b2</b>), (<b>b4</b>) the HRTEM image of the α″ and the BCC matrix; (<b>c1</b>) the HRTEM image of the ω phase and (<b>c2</b>) the responding FFT; and (<b>c3</b>) the HRTEM image of the B2 phase and (<b>c4</b>) the responding FFT.</p>
Full article ">Figure 7
<p>The mechanical properties of Al<sub>x</sub> MEAs at room temperature. (<b>a</b>) Tensile curves at room temperature; (<b>b</b>) the compression curves of Al20; and the corresponding work-hardening rate curves of (<b>c</b>) Al12 and (<b>d</b>) Al20.</p>
Full article ">Figure 8
<p>The fracture morphologies of the Al<sub>x</sub> MEAs. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20.</p>
Full article ">Figure 9
<p>The DSC curves of the Al12 and Al20 MEAs.</p>
Full article ">Figure 10
<p>The XRD patterns of Al20 before and after the tensile test.</p>
Full article ">
35 pages, 14744 KiB  
Review
Review of the Properties of GaN, InN, and Their Alloys Obtained in Cubic Phase on MgO Substrates by Plasma-Enhanced Molecular Beam Epitaxy
by Edgar López Luna and Miguel Ángel Vidal
Crystals 2024, 14(9), 801; https://doi.org/10.3390/cryst14090801 - 11 Sep 2024
Viewed by 933
Abstract
Gallium nitride (GaN) semiconductors and their broadband InGaN alloys in their hexagonal phase have been extensively studied over the past 30 years and have allowed the development of blue-ray lasers, which are essential disruptive developments. In addition to high-efficiency white light-emitting diodes, which [...] Read more.
Gallium nitride (GaN) semiconductors and their broadband InGaN alloys in their hexagonal phase have been extensively studied over the past 30 years and have allowed the development of blue-ray lasers, which are essential disruptive developments. In addition to high-efficiency white light-emitting diodes, which have revolutionized lighting technologies and generated a great industry around these semiconductors, several transistors have been developed that take advantage of the characteristics of these semiconductors. These include power transistors for high-frequency applications and high-power transistors for power electronics, among other devices, which have far superior achievements. However, less effort has been devoted to studying GaN and InGaN alloys grown in the cubic phase. The metastable or cubic phase of III-N alloys has superior characteristics compared to the hexagonal phase, mainly because of the excellent symmetry. It can be used to improve lighting technologies and develop other devices. Indium gallium nitride, InxGa1−xN alloy, has a variable band interval of 0.7 to 3.4 eV that covers almost the entire solar spectrum, making it a suitable material for increasing the efficiencies of photovoltaic devices. In this study, we successfully synthesized high-quality cubic InGaN films on MgO (100) substrates using plasma-assisted molecular beam epitaxy (PAMBE), demonstrating tunable emissions across the visible spectrum by varying the indium concentration. We significantly reduced the defect density and enhanced the crystalline quality by using an intermediate cubic GaN buffer layer. We not only developed a heterostructure with four GaN/InGaN/GaN quantum wells, achieving violet, blue, yellow, and red emissions, but also highlighted the immense potential of cubic InGaN films for high-efficiency light-emitting diodes and photovoltaic devices. Achieving better p-type doping levels is crucial for realizing diodes with excellent performance, and our findings will pave the way for this advancement. Full article
(This article belongs to the Special Issue Reviews of Crystal Engineering)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction patterns of the GaN films grown with different Ga cell temperatures T<sub>Ga</sub>. The patterns are vertically offset for clarity.</p>
Full article ">Figure 2
<p>X-ray pole figures of GaN films corresponding to (0002) hexagonal plane growth with T<sub>Ga</sub> of (<b>a</b>) 900 °C, (<b>b</b>) 940 °C, (<b>c</b>) 955 °C, (<b>d</b>) 970 °C and corresponding to (002) cubic plane with T<sub>Ga</sub> of (<b>e</b>) 900 °C, (<b>f</b>) 940 °C, (<b>g</b>) 955 °C, (<b>h</b>) 970 °C.</p>
Full article ">Figure 3
<p>X-ray pole figures of GaN films growth with T<sub>Ga</sub> = 940 °C for (<b>a</b>) (10<span class="html-overline">1</span>0) and (<b>b</b>) (101<span class="html-overline">1</span>) hexagonal planes.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM surface micrograph of GaN film growth at T<sub>Ga</sub> = 940 °C. (<b>b</b>) Streaky RHEED pattern for the MgO surface, indicating a two-dimensional surface morphology. (<b>c</b>) GaN RHEED pattern of [100] zincblende azimuth, at the start of the growth on MgO (100) substrate. (<b>d</b>) RHEED pattern of GaN wurtzite azimuth [11-20] with growth aligned with [110] direction of MgO (100) substrate. A model of the growth at the interface of GaN over MgO is shown in (<b>e</b>). Cubic GaN domains oriented in the [100] and [010] directions, and (<b>f</b>) shows the [11-20] azimuth of the c-plane, growing perpendicular to the film surface of GaN, aligned on the [110] direction of the MgO (100) surface.</p>
Full article ">Figure 5
<p>SEM figure (<b>a</b>) and high-resolution cross-sectional TEM images (<b>b</b>,<b>c</b>) of the film grown at TGa = 970 °C.</p>
Full article ">Figure 6
<p>(<b>a</b>) 2θ-ω XRD scan of c-InGaN/GaN samples with different In mole fraction using MgO (100) substrates. (<b>b</b>) Lattice parameter vs. In molar fraction [<a href="#B70-crystals-14-00801" class="html-bibr">70</a>].</p>
Full article ">Figure 7
<p>Variation of (αhν)<sup>2</sup> vs. photon energy (hν) with In molar fraction of c-InGaN [<a href="#B60-crystals-14-00801" class="html-bibr">60</a>]. Reprinted from Publication Journal of Crystal Growth 418 (2015) 120–125, V.D. Compeán García, I.E. Orozco Hinostroza, A. Escobosa Echavarría, E. López Luna, A.G. Rodríguez, M.A. Vidal, Bulk lattice parameter and band gap of cubic In X Ga1−X N (001) alloys on MgO (100) substrates Copyright (2015), with permission from Elsevier.</p>
Full article ">Figure 8
<p>The compositional bandgap variation vs. In is found to exhibit a bowing parameter in a quadratic form according to the expression in Equation (1). Reprinted from Publication Journal of Crystal Growth 418 (2015) 120–125, V.D. Compeán García, I.E. Orozco Hinostroza, A. Escobosa Echavarría,E. López Luna, A.G. Rodríguez, M.A. Vidal, Bulk lattice parameter and band gap of cubic In X Ga1−X N (001) alloys on MgO (100) substrates Copyright (2015), with permission from Elsevier. Ref. [<a href="#B60-crystals-14-00801" class="html-bibr">60</a>].</p>
Full article ">Figure 9
<p>Scheme of two hexagonal InN domains with 30° plane rotation.</p>
Full article ">Figure 10
<p>(<b>a</b>) Pole figure of InN taken in the [01-10] plane showing 12-fold symmetry and (<b>b</b>) XRD curve of InN growth on LT-InN/MgO (100).</p>
Full article ">Figure 11
<p>SEM images with their corresponding RHEED pattern (insets) for samples of (<b>a</b>) GaN/MgO, the buffer layer prior to the InN growth, (<b>b</b>) InN at T<sub>In</sub> = 750 °C, (<b>c</b>) InN at T<sub>In</sub> = 790 °C, and (<b>d</b>) InN at T<sub>In</sub> = 840 °C [<a href="#B95-crystals-14-00801" class="html-bibr">95</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) XRD curve of c-InN growth on c-GaN/MgO (100) layer under metal-rich conditions. and (<b>b</b>) pole figure of c-InN/c-GaN/MgO (100) taken in the (0002)/(111) plane showing 4-fold symmetry corresponding to c-InN (111).</p>
Full article ">Figure 13
<p>Surface lattice constant evolution as a function of the deposited monolayers of GaN on MgO (100).</p>
Full article ">Figure 14
<p>Surface lattice constant evolution as a function of the deposited monolayers of In<sub>x</sub>Ga<sub>1−x</sub>N on c-GaN/MgO (100).</p>
Full article ">Figure 15
<p>Surface lattice constant evolution as a function of the deposited monolayers of GaN on MgO (100) [<a href="#B83-crystals-14-00801" class="html-bibr">83</a>].</p>
Full article ">Figure 16
<p>Critical layer thickness in nm plotted over the indium molar fraction. The black solid line is the Fisher fit. Black star are measurements from Ref. [<a href="#B94-crystals-14-00801" class="html-bibr">94</a>], the blue square is obtained from Ref. [<a href="#B113-crystals-14-00801" class="html-bibr">113</a>], and the red squares are the data points from Ref. [<a href="#B107-crystals-14-00801" class="html-bibr">107</a>].</p>
Full article ">Figure 17
<p>Photoluminescence of individual InGaN QWs with In concentrations of x = 0.10 (<b>a</b>), x = 0.40 (<b>b</b>), and x = 0.47 (<b>c</b>). (<b>d</b>) Schematic diagram of the QWs, from [<a href="#B114-crystals-14-00801" class="html-bibr">114</a>].</p>
Full article ">Figure 18
<p>Bandgap of InGaN vs. In molar fraction and quantum wells for different thicknesses (3, 4, and 10 nm) and bandgap taken from <a href="#crystals-14-00801-f008" class="html-fig">Figure 8</a> [<a href="#B60-crystals-14-00801" class="html-bibr">60</a>].</p>
Full article ">Figure 19
<p>Photoluminescence spectra of the structure with 4 active layers at 10 K. The fitting of 5 peaks is shown, including the GaN buffer layer at 3.26 eV. The colored lines represent the corresponding fits: the dark blue line corresponds to the GaN buffer layer, while the red, light blue, green, and brown lines represent the emission from the active layers.</p>
Full article ">Figure 20
<p>The surface topography of different doping regions by atomic force microscopy in (<b>a</b>) topography, (<b>b</b>) CPD map of unintentional n-type c-GaN, (<b>c</b>) CPD measurement, and (<b>d</b>) optical image of the terrace bilayer of the unintentional c-GaN n-type and the p-type c-GaN: Mg layer, figure (<b>e</b>) shows the heterostructure of a p-n junction −/+, and finally, (<b>f</b>) the topography and (<b>g</b>) CPD map of the p-type c-GaN: Mg.</p>
Full article ">Figure 21
<p>Electrical characterization current density vs. voltage (J–V) of c-GaN/c-GaN-p homojunction.</p>
Full article ">
14 pages, 5290 KiB  
Article
Influence of Solid-Phase and Melt-Quenching Na3Fe2(PO4)3 Polycrystal Production Technology on Their Structure and Ionic Conductivity
by A. S. Nogai, A. A. Nogai, D. E. Uskenbaev, E. A. Nogai, A. B. Utegulov, P. A. Dunayev, A. S. Tolegenova, Bazarbek Assyl-Dastan Bazarbekuly and A. A. Abikenova
J. Compos. Sci. 2024, 8(9), 354; https://doi.org/10.3390/jcs8090354 - 9 Sep 2024
Viewed by 415
Abstract
This article studies the influence of solid-phase (type 1 samples) and melt-quenching (type 2 samples) technological modes of obtaining Na3Fe2(PO4)3 polycrystals on their structures and ion-conducting properties. α-Na3Fe2(PO4)3 polycrystals [...] Read more.
This article studies the influence of solid-phase (type 1 samples) and melt-quenching (type 2 samples) technological modes of obtaining Na3Fe2(PO4)3 polycrystals on their structures and ion-conducting properties. α-Na3Fe2(PO4)3 polycrystals of the 1st type are formed predominantly under an isothermal firing regime, and the synthesis of the 2nd type is carried out under sharp temperature gradient conditions, contributing to the formation of glassy precursors possessing a reactive and deformed structure, in which the crystallization of crystallites occurs faster than in precursors obtained under isothermal firing. The elemental composition of α-Na3Fe2(PO4)3 type 2 polycrystals is maintained within the normal range despite the sharp non-equilibrium thermodynamic conditions of synthesis. The microstructure of the type 1 Na3Fe2(PO4)3 polycrystals is dominated by chaotically arranged crystallites of medium (7–10 μm) and large (15–35 μm) sizes, while the polycrystals of type 2 are characterized by the preferential formation of small (3–4 μm) and medium (7–10 μm) crystallites, causing uniaxial deformations in their structure, which contribute to a partial increase in their symmetry. The advantage of type 2 polycrystals is that they have higher density and conductivity and are synthesized faster than type 1 samples by a factor of 4. The article also considers the issues of crystallization in a solid-phase precursor from the classical point of view, i.e., the process of the formation of small solid-phase nuclei in the metastable phase and their growth to large particles due to association with small crystallites using phase transitions. Possible variants and models of crystallite growth in Na3Fe2(PO4)3 polycrystals, as well as distinctive features of crystallization between two types of samples, are discussed. Full article
(This article belongs to the Section Composites Manufacturing and Processing)
Show Figures

Figure 1

Figure 1
<p>X-ray diffractograms of (<b>a</b>) amorphous phases of Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> (after firing at 350 °C for 2 h) and (<b>b</b>) type 1 amorphous–crystalline precursors (after one firing at 600 °C for 7 h) (The designations of the crystallographic planes [hkl] are given on the diffraction peaks).</p>
Full article ">Figure 2
<p>X-ray diffractograms of type 2 glassy precursors prepared by melting amorphous phases of Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> at 950 °C and cooling the melt: (<b>a</b>) from thin plates; (<b>b</b>) thick plates.</p>
Full article ">Figure 3
<p>X-ray diffractogram of Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> polycrystals: type 1 and type 2.</p>
Full article ">Figure 4
<p>Microstructures of α-Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> polycrystals of type 1 (<b>a</b>) and type 2 (<b>b</b>).</p>
Full article ">Figure 5
<p>Dependence of the fraction of crystallites on their sizes in α-Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> polycrystals: for type 1 samples and for type 2 samples.</p>
Full article ">Figure 6
<p>Temperature dependence of the Gibbs energy (for a crystallite with G<sub>S</sub>(T) energy) and (for a metastable phase with G<sub>MS</sub>(T) energy). Crystallization processes: (<b>a</b>) the arrow indicates the transition from the liquid phase (G<sub>L</sub>) to the solid phase (G<sub>S</sub>), the dotted line indicates the temperature of the phase transition (T<sub>C</sub>); (<b>b</b>) the arrow indicates the transition from the metastable phase (∆G<sub>MS</sub>) to the solid phase. The point at which the G<sub>Snk</sub> nucleus forms corresponds to the PT temperature T<sub>c</sub>.</p>
Full article ">Figure 7
<p>Schematic representation of the FP process from the amorphous phase to crystalline G<sub>MS</sub> → G<sub>S</sub> due to ∆G<sub>MS</sub> energy: (<b>a</b>) for type 1 samples; (<b>b</b>) for type 2 samples. Moreover, ∆G<sub>MS2</sub> ˃ ∆G<sub>MS1</sub>.</p>
Full article ">Figure 8
<p>Elemental composition of Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> polycrystals: (<b>a</b>) type 1 and (<b>b</b>) type 2.</p>
Full article ">Figure 9
<p>Temperature dependences of the ionic conductivity of Na<sub>3</sub>Fe<sub>2</sub>(PO<sub>4</sub>)<sub>3</sub> for the type 1 (black line) and type 2 polycrystals (red line), α-, β-, and γ-phases are separated from each other by dotted lines.</p>
Full article ">
Back to TopTop