[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (419)

Search Parameters:
Keywords = neutron diffraction

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 11321 KiB  
Article
Uncovering the Mechanisms of Long-Range Magnetic Order in [Mn(mal)(H2O)]n: Insights from Microscopic and Macroscopic Magnetic Analysis
by Fernando S. Delgado, Laura Cañadillas-Delgado, Juan Rodríguez-Carvajal, Óscar Fabelo and Jorge Pasán
Magnetochemistry 2024, 10(12), 109; https://doi.org/10.3390/magnetochemistry10120109 - 20 Dec 2024
Viewed by 385
Abstract
In this study, we investigate the magnetic properties of the molecular compound [Mn(mal)(H2O)]ₙ (mal = dianion of malonic acid) by integrating microscopic and macroscopic characterization, combining unpolarized neutron diffraction and magnetometry measurements. Neutron diffraction, though non-commonly applied to molecular compounds, proved [...] Read more.
In this study, we investigate the magnetic properties of the molecular compound [Mn(mal)(H2O)]ₙ (mal = dianion of malonic acid) by integrating microscopic and macroscopic characterization, combining unpolarized neutron diffraction and magnetometry measurements. Neutron diffraction, though non-commonly applied to molecular compounds, proved essential for fully resolving the magnetic structure, as well as overcoming challenges such as hydrogen-related incoherent scattering and difficulties in accurately locating light atoms. Our neutron data provided critical structural details, including the precise location of hydrogen atoms, especially those associated with crystallization water molecules. By conducting low-temperature measurements below the magnetic ordering temperature, we identified the correct Shubnikov space group (Pc’a21’) and established a magnetic model consistent with the observed weak ferromagnetism. Our findings reveal that the compound presents a spin-canted structure with a weak ferromagnetic signal along the b-axis. This signal originates primarily from antisymmetric exchange interactions rather than single-ion anisotropy, consistent with the isotropic nature of the Mn(II) (6A1g) ground state. The combined neutron diffraction and magnetometry results provide a comprehensive understanding of how structural and symmetry factors influence the magnetic properties of malonate-based manganese compounds. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Detailed view of the coordination mode of the malonate ligand in [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub>. The coordination environment of the manganese atom in the asymmetric unit has been completed to enhance understanding of the structure. Atoms in the asymmetric unit are represented as ellipsoids with a 50% probability. (<b>b</b>) Coordination environment around Mn with blue-dashed lines indicates hydrogen bonds according to the neutron-refined model at 5 K. Symmetry operators: (a) = −1/2 + <span class="html-italic">x</span>, 1 − <span class="html-italic">y</span>, <span class="html-italic">z</span>; (b) = −1/2 + <span class="html-italic">x</span>, −<span class="html-italic">y</span>, <span class="html-italic">z</span> (c) = 2 − <span class="html-italic">x</span>, 1 − <span class="html-italic">y</span>, −1/2 + <span class="html-italic">z</span>; (d) = 2 − <span class="html-italic">x</span>, −<span class="html-italic">y</span>, −1/2 + <span class="html-italic">z</span>; (e) = 5/2−<span class="html-italic">x</span>, <span class="html-italic">y</span>, 1/2 + <span class="html-italic">z</span>.</p>
Full article ">Figure 2
<p>(<b>a</b>) View along the <span class="html-italic">c</span>-axis and (<b>b</b>) view along the <span class="html-italic">a</span>-axis of the refined structural model for [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub> at 5 K, obtained using neutron powder diffraction data from the G4.2 diffractometer. For clarity, alternating layers stacked along the <span class="html-italic">c</span>-axis are shown in red and green to illustrate the <span class="html-italic">ABAB</span> stacking sequence.</p>
Full article ">Figure 3
<p>(<b>Top</b>) Temperature dependence of the magnetization for [Mn(mal)(H<sub>2</sub>O)<sub>2</sub>]<sub>n</sub>, showing zero-field-cooled (open triangles) and field-cooled (solid circles) measurements. Solid lines are included as a visual guide to aid in identifying trends. (<b>Bottom</b>) Temperature dependence of out-of-phase <span class="html-italic">ac</span> susceptibility, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>χ</mi> </mrow> <mrow> <mi mathvariant="normal">M</mi> </mrow> </msub> </mrow> </semantics></math>, measured under an oscillating field of 1 G at a frequency of 333 Hz.</p>
Full article ">Figure 4
<p>(<b>a</b>) View of the Fourier map around O(1w), derived from low-temperature (5 K) neutron diffraction data, using the X-ray model (without hydrogen atoms of the crystallization water) as the starting point. (<b>b</b>) Fourier map around O(2w), following the placement of hydrogen atoms on O(1w).</p>
Full article ">Figure 5
<p>Rietveld refinement of [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub> collected at 5 K, measured on the G4.2 diffractometer. Experimental data and fitted profiles are represented by red and black lines, respectively, while the blue line indicates the residuals of the refinement. Vertical green ticks mark the positions of the Bragg reflections. The gap between 143° and 151° corresponds to an excluded region where contributions from the sample environment become visible. The final Bragg R-factor obtained for the paramagnetic phase in the last refinement cycle is 9.2%.</p>
Full article ">Figure 6
<p>(<b>a</b>) Thermodiffractogram of [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub> collected on the G4.1 instrument. (<b>b</b>) Thermodiffractogram of the magnetic signal, calculated as the difference between the pattern measured at different temperatures and the one measured at 5 K in the paramagnetic phase (difference pattern). The Néel temperature is highlighted with a black-dashed line in both figures.</p>
Full article ">Figure 7
<p>Magnetic structure of [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub> compound in the <span class="html-italic">Pc’a</span>2<sub>1</sub>’ magnetic space group. (<b>Top</b>) View along the <span class="html-italic">c</span>-axis of one of the layers of [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub> extended in the <span class="html-italic">ab</span>-plane. The unit cell is indicated in blue for reference. The main magnetic interactions between Mn ions via <span class="html-italic">anti</span>–<span class="html-italic">anti</span> exchange (<span class="html-italic">J</span>) are shown with a green bond, while the secondary interaction (<span class="html-italic">j</span>) is depicted with an orange-dashed line. The magnetic moments are represented in blue. (<b>Bottom</b>) Detailed view in the [011] direction of the magnetic chain formed by Mn atoms linked through carboxylate bridges in an <span class="html-italic">anti</span>–<span class="html-italic">anti</span> configuration.</p>
Full article ">Figure 8
<p>Rietveld refinement of the difference pattern (1.5–5 K) for [Mn(mal)(H<sub>2</sub>O)]n, measured on the G4.1 diffractometer, was performed using the <span class="html-italic">Pc’a</span>2<sub>1</sub>’ Shubnikov space group (Γ4), as described in the main text. The refinement corresponds to a model where <span class="html-italic">m<sub>x</sub></span> is fixed to zero and <span class="html-italic">m<sub>y</sub></span> is fixed to 0.05 μ<sub>B</sub> (for further details, see the main text). The resulting Bragg R-factor for the magnetic reflections is 10.6%. Experimental data and the fitted profiles are shown by the red and black lines, respectively, with the blue line representing the residuals of the refinement. Red vertical ticks mark pure magnetic reflections, while blue ticks denote both magnetic and nuclear Bragg peaks (note that the nuclear contribution is zero due to the difference pattern).</p>
Full article ">Figure 9
<p>(<b>a</b>) Refined magnetic structure of [Mn(mal)(H<sub>2</sub>O)]<sub>n</sub>, with magnetic moments depicted as blue arrows. The nuclear structure is shown in transparent mode for clarity. (<b>b</b>) View along the <span class="html-italic">a</span>-axis, where magnetic moment components along the <span class="html-italic">b</span>-axis (representing the weak ferromagnetic component—see main text) are magnified by a factor of 10 for visibility. (<b>c</b>) View along the <span class="html-italic">c</span>-axis, with the <span class="html-italic">m<sub>y</sub></span> component similarly enhanced as in (<b>b</b>). Different stacking layers along the <span class="html-italic">c</span>-axis are shown with varying transparencies, and magnetic moments in each layer are represented in alternating blue and green hues to improve visual distinction.</p>
Full article ">Figure 10
<p>Evolution of the magnetic moment as a function of temperature. The solid line represents the power law fit of the data in the critical region (see main text).</p>
Full article ">
29 pages, 6836 KiB  
Review
Advanced Characterization of Solid-State Battery Materials Using Neutron Scattering Techniques
by Eric Novak, Luke Daemen and Niina Jalarvo
Materials 2024, 17(24), 6209; https://doi.org/10.3390/ma17246209 - 19 Dec 2024
Viewed by 279
Abstract
Advanced batteries require advanced characterization techniques, and neutron scattering is one of the most powerful experimental methods available for studying next-generation battery materials. Neutron scattering offers a non-destructive method to probe the complex structural and chemical processes occurring in batteries during operation in [...] Read more.
Advanced batteries require advanced characterization techniques, and neutron scattering is one of the most powerful experimental methods available for studying next-generation battery materials. Neutron scattering offers a non-destructive method to probe the complex structural and chemical processes occurring in batteries during operation in truly in situ/in operando measurements with a high sensitivity to battery-relevant elements such as lithium. Neutrons have energies comparable to the energies of excitations in materials and wavelengths comparable to atomic distances in the solid state, thus giving access to study structural and dynamical properties of materials on an atomic scale. In this review, a broad overview of selected neutron scattering techniques is presented to illustrate how neutron scattering can be used to gain invaluable information of solid-state battery materials, with a focus on in situ/in operando methods. These techniques span multiple decades of length and time scales to uncover the complex processes taking place fundamentally on the atomic scale and to determine how these processes impact the macroscale properties and performance of functional battery systems. This review serves the solid-state battery research community by examining how the unique capabilities of neutron scattering can be applied to answer critical and unresolved questions of materials research in this field. A thorough and broad perspective is provided with numerous practical examples showing these techniques in action for battery research. Full article
(This article belongs to the Special Issue Local Structure Characterization for Complex Functional Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Neutron diffraction patterns of a LiNiO<sub>2</sub> | | graphite full cell during the first cycle [<a href="#B37-materials-17-06209" class="html-bibr">37</a>]. (<b>b</b>) Voltage profile as a function of time. Evolution of the (<b>c</b>) LiNiO2 and (<b>d</b>) graphite Bragg peaks, (<b>e</b>) lattice parameters, and (<b>f</b>) unit cell volume during cycling as a function of Li concentration in Li<sub>x</sub>NiO<sub>2</sub>. Reprinted with permission from [<a href="#B37-materials-17-06209" class="html-bibr">37</a>]. Copyright 2021 John Wiley and Sons.</p>
Full article ">Figure 2
<p>(<b>a</b>) Real space information that can be obtained from <span class="html-italic">G</span>(<span class="html-italic">r</span>) [<a href="#B47-materials-17-06209" class="html-bibr">47</a>]. (<b>b</b>) Comparison of the atomic structure of graphdiyne to <span class="html-italic">G</span>(<span class="html-italic">r</span>) showing how the peaks are a real-space representation of atomic distances compared to an atom located at the origin [<a href="#B47-materials-17-06209" class="html-bibr">47</a>,<a href="#B48-materials-17-06209" class="html-bibr">48</a>]. Reprinted with permission from [<a href="#B47-materials-17-06209" class="html-bibr">47</a>]. Copyright 2016 Springer Nature. Reprinted with permission from [<a href="#B48-materials-17-06209" class="html-bibr">48</a>]. Copyright 2018 John Wiley and Sons.</p>
Full article ">Figure 3
<p>(<b>a</b>) <span class="html-italic">G</span>(<span class="html-italic">r</span>) for annealed (A48 and A 240) and non-annealed (FC and SC) LiNi<sub>0.5</sub>Mn<sub>1.5</sub>O<sub>4</sub> samples [<a href="#B50-materials-17-06209" class="html-bibr">50</a>]. The local structure is virtually identical for all 4 samples, but differences are observed at longer length scales corresponding to varying degrees of disorder. (<b>b</b>) Potential local vs. long-range ordering of the Ni/Mn sites to produce either an ordered or disordered global structure. Reprinted with permission from [<a href="#B50-materials-17-06209" class="html-bibr">50</a>]. Copyright 2016 American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>a</b>) Simulated crystal structure of Li<sub>2.94</sub>PO<sub>3.5</sub>N<sub>0.31</sub> generated from ab initio molecular dynamics. O, N, Li, and P atoms are colored red, blue, green, and gray, respectively [<a href="#B54-materials-17-06209" class="html-bibr">54</a>]. The apical and double-bridging N configurations are shown. (<b>b</b>) Comparison of the experimental and simulated PDF. Reprinted with permission from [<a href="#B54-materials-17-06209" class="html-bibr">54</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) SANS curves normalized by <span class="html-italic">Q<sup>p</sup></span> for the pristine LiC<sub>x</sub>-air system to highlight changes in the intergrain nanopores [<a href="#B62-materials-17-06209" class="html-bibr">62</a>]. (<b>b</b>) SANS curves for graphite, LiC<sub>6</sub>. And LiC<sub>6</sub> + dEC. Inset has the scattering law exponents, showing that the SEI roughens the graphite surfaces. Reprinted with permission from [<a href="#B62-materials-17-06209" class="html-bibr">62</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic view of the cell used for in operando SANS measurements [<a href="#B64-materials-17-06209" class="html-bibr">64</a>]. (<b>b</b>) Scattering length densities for the cell components and major SEI components (electrolyte reduction products). Dashed lines represent the contrast difference between the two electrolytes. SANS intensity as a function of first discharge time, momentum transfer, and cell voltage for (<b>c</b>) 1 M LiTFSi/PC and (<b>d</b>) 4 M LiTFSi/PC electrolyte solutions. Reprinted with permission from [<a href="#B64-materials-17-06209" class="html-bibr">64</a>]. Copyright 2019 RSC.</p>
Full article ">Figure 7
<p>Scattering length density as a function of distance from the substrate [<a href="#B72-materials-17-06209" class="html-bibr">72</a>]. Reprinted with permission from [<a href="#B72-materials-17-06209" class="html-bibr">72</a>]. Copyright 2021 IOP.</p>
Full article ">Figure 8
<p>(<b>a</b>) Neutron reflectometry data for a NiO/LiPON solid-state battery for the (blue) initial pristine state, (red) after Li plating, and (cyan) after Li stripping [<a href="#B73-materials-17-06209" class="html-bibr">73</a>]. The data are plotted as <span class="html-italic">RQ</span><sup>4</sup> to enhance the profile features. The black lines are a fit to a multi-layer thin film structure. (<b>b</b>) Scattering length density profiles (SLD) calculated from fits to the reflectivity data with a diagram showing the thickness of the different layers. Reprinted with permission from [<a href="#B73-materials-17-06209" class="html-bibr">73</a>]. Copyright 2023 American Chemical Society.</p>
Full article ">Figure 9
<p>Evolution of the Li distribution in the battery at a series of charge/discharge stages. High attenuation due to the presence of <sup>6</sup>Li is colored white. Dendrite formation is observed during charging and disappears upon discharge [<a href="#B86-materials-17-06209" class="html-bibr">86</a>]. Reprinted with permission from [<a href="#B86-materials-17-06209" class="html-bibr">86</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 10
<p>Neutron tomography for solid-state sulfur cathodes. (<b>a</b>) 3D reconstruction of the discharged (Li in cathode) and charged states, along with the difference showing the location of the mobile Li shown in green. Neutron imaging slices of 39 μm thick slices of the cathode showing the lithium distribution for the (<b>b</b>) discharged and (<b>c</b>) charged states [<a href="#B93-materials-17-06209" class="html-bibr">93</a>]. Reprinted with permission from [<a href="#B93-materials-17-06209" class="html-bibr">93</a>]. Copyright 2023 John Wiley and Sons. (<b>d</b>) Total attenuation change for the discharged (orange) and recharged (blue) states. The difference (green) shows the mobile Li.</p>
Full article ">Figure 11
<p>(<b>a</b>) X-ray and (<b>b</b>) neutron imaging of commercial Li/MnO<sub>2</sub> cells that contain a wound electrode/separator/current collector assembly. The X-ray images are sensitive to the highly attenuating Ni current collectors, while the neutron technique is sensitive to the Li distributions and electrolyte. Cracks and delamination can be observed by X-rays in the cathode as the cell expands upon lithium insertion. Different colored arrows highlight regions of interest with features explained in the correspondingly colored text boxes. Virtual unfolding of the separator/electrode assemblies using (<b>c</b>) X-rays and (<b>d</b>) neutrons [<a href="#B88-materials-17-06209" class="html-bibr">88</a>]. Reprinted with permission from [<a href="#B88-materials-17-06209" class="html-bibr">88</a>]. Copyright 2020 Springer Nature.</p>
Full article ">Figure 12
<p>(<b>a</b>) INS spectra of Li<sub>3</sub>InCl<sub>6</sub> measured at 5 K at the VISION spectrometer [<a href="#B100-materials-17-06209" class="html-bibr">100</a>]. (<b>b</b>) INS spectra for EC, LiC<sub>6</sub>, LiC<sub>6</sub> + EC/DMC, and washed LiC<sub>6</sub> + EC/DMC [<a href="#B62-materials-17-06209" class="html-bibr">62</a>]. The asterisks (*) correspond to EC peaks, while daggers (†) are associated with PEO-type peaks. Reprinted with permission from [<a href="#B62-materials-17-06209" class="html-bibr">62</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 13
<p>INS phonon density of states for (<b>a</b>) Li<sub>3</sub>PO<sub>4</sub>, (<b>b</b>) Li<sub>3</sub>PS<sub>4</sub>, (<b>c</b>) Li<sub>3.4</sub>Ge<sub>0.4</sub>P<sub>0.6</sub>O<sub>4</sub>, and (<b>d</b>) Li<sub>3.25</sub>Ge<sub>0.25</sub>P<sub>0.75</sub>S<sub>4</sub>. The top line are the experimental data collected at 100 K, the bottom line is calculated computationally at 0 K, and the shaded region is the Li phonon contributions. (<b>e</b>) A comparison of the oxidation voltage as a function of anion phonon band center [<a href="#B99-materials-17-06209" class="html-bibr">99</a>]. Reprinted with permission from [<a href="#B99-materials-17-06209" class="html-bibr">99</a>]. Copyright 2018 RSC.</p>
Full article ">Figure 14
<p>(<b>a</b>) Structural features of NaPS<sub>4</sub> [<a href="#B133-materials-17-06209" class="html-bibr">133</a>], (<b>b</b>) ionic conductivity of NaPS<sub>4</sub> in both the β and γ phases [<a href="#B133-materials-17-06209" class="html-bibr">133</a>], (<b>c</b>) elastic scan of Na<sub>3</sub>PS<sub>4</sub>, and (<b>d</b>) QENS spectra of Na<sub>3</sub>PS<sub>4</sub> at temperatures from 100 to 690 °C illustrating different dynamical behavior for three different structural phases [<a href="#B135-materials-17-06209" class="html-bibr">135</a>]. (<b>a</b>,<b>b</b>) Reprinted with permission from [<a href="#B133-materials-17-06209" class="html-bibr">133</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">
14 pages, 3148 KiB  
Article
Liquid Structure of Magnesium Aluminates
by Viviana Cristiglio, Irina Pozdnyakova, Aleksei Bytchkov, Gabriel J. Cuello, Sandro Jahn, Didier Zanghi, Séverine Brassamin, James W. E. Drewitt and Louis Hennet
Materials 2024, 17(24), 6173; https://doi.org/10.3390/ma17246173 - 17 Dec 2024
Viewed by 367
Abstract
Magnesium aluminates (MgO)x(Al2O3)1−x belong to a class of refractory materials with important applications in glass and glass–ceramic technologies. Typically, these materials are fabricated from high-temperature molten phases. However, due to the difficulties in making measurements [...] Read more.
Magnesium aluminates (MgO)x(Al2O3)1−x belong to a class of refractory materials with important applications in glass and glass–ceramic technologies. Typically, these materials are fabricated from high-temperature molten phases. However, due to the difficulties in making measurements at very high temperatures, information on liquid-state structure and properties is limited. In this work, we employed the method of aerodynamic levitation with CO2 laser heating at large scale facilities to study the structure of liquid magnesium aluminates in the system (MgO)x(Al2O3)1−x, with x = 0.33, 0.5, and 0.75, using X-ray and neutron diffraction. We determined the structure factors and corresponding pair distribution functions, providing detailed information on the short-range structural order in the liquid state. The local structures were similar across the range of compositions studied, with average coordination numbers of n¯AlO4.5  and n¯MgO5.1 and interatomic distances of rAlO=1.761.78 Å and rMgO=1.931.95 Å. The results are in good agreement with previous molecular dynamics simulations. For the spinel endmember MgAl2O4 (x = 0.5), the average Mg-O and Al-O coordination numbers gave rise to conflicting values for the inversion coefficient χ, indicating that the structural formula used to describe the solid-state order-disorder transition is not applicable in the liquid state. Full article
(This article belongs to the Section Materials Physics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The total structure factors (<b>a</b>) <span class="html-italic">S<sup>X</sup>(Q)</span> and (<b>b</b>) <span class="html-italic">S<sup>N</sup>(Q)</span> for the (MgO)<span class="html-italic"><sub>x</sub></span>(Al<sub>2</sub>O<sub>3</sub>)<sub>1−<span class="html-italic">x</span></sub> liquids as measured using X-ray and neutron diffraction, respectively. The MA and M3A data are displaced vertically in increments of 0.5 for clarity. For comparison with the X-ray data, the <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>S</mi> </mrow> <mrow> <mi>N</mi> </mrow> </msup> <mfenced separators="|"> <mrow> <mi>Q</mi> </mrow> </mfenced> </mrow> </semantics></math> is plotted up to 16 Å<sup>−1</sup>.</p>
Full article ">Figure 2
<p>The total pair distribution functions <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>G</mi> </mrow> <mrow> <mi>X</mi> </mrow> </msup> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> <mo> </mo> </mrow> </semantics></math>(<b>a</b>) and <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>G</mi> </mrow> <mrow> <mi>N</mi> </mrow> </msup> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math> (<b>b</b>) for the (MgO)<span class="html-italic"><sub>x</sub></span>(Al<sub>2</sub>O<sub>3</sub>)<sub>1−<span class="html-italic">x</span></sub> liquids as obtained by Fourier transforming the corresponding total structure factors shown in <a href="#materials-17-06173-f001" class="html-fig">Figure 1</a>. The MA and M3A data are displaced vertically in increments of 0.5 for clarity.</p>
Full article ">Figure 3
<p>(<b>a</b>) The total structure factors <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>S</mi> </mrow> <mrow> <mi>X</mi> </mrow> </msup> <mfenced separators="|"> <mrow> <mi>Q</mi> </mrow> </mfenced> </mrow> </semantics></math> and (<b>b</b>) pair distribution functions <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>G</mi> </mrow> <mrow> <mi>X</mi> </mrow> </msup> <mfenced separators="|"> <mrow> <mi>r</mi> </mrow> </mfenced> </mrow> </semantics></math> for liquid MgAl<sub>2</sub>O<sub>4</sub> at various temperatures in the liquid and supercooled states. The data are displaced vertically in steps of 0.25 for clarity.</p>
Full article ">Figure 4
<p>Al-O (<b>a</b>) and Mg-O (<b>b</b>) coordination number distribution as function of composition.</p>
Full article ">Figure 5
<p>Gaussian fit to the X-ray total correlation function <span class="html-italic">T<sup>X</sup>(r)</span> of the equimolar composition MgAl<sub>2</sub>O<sub>4</sub>. The inset shows the decomposition into 2 Gaussians corresponding to Al-O and Mg-O distances.</p>
Full article ">Figure 6
<p>Gaussian fit to the neutron total correlation function <span class="html-italic">T<sup>N</sup>(r)</span> for the equimolar composition MgAl<sub>2</sub>O<sub>4</sub>. The inset shows the decomposition into 2 Gaussians corresponding to Al-O and Mg-O distances.</p>
Full article ">
21 pages, 6488 KiB  
Article
X-Ray Diffraction Line Broadening of Irradiated Zr-2.5Nb Alloys
by Malcolm Griffiths
Metals 2024, 14(12), 1446; https://doi.org/10.3390/met14121446 - 17 Dec 2024
Viewed by 281
Abstract
The evolution of the mechanical properties of Zr-2.5Nb pressure tubing during irradiation is dependent on dislocation loop densities that are represented by the broadening of X-ray diffraction lines. Empirical models for the integral breadth of the diffraction peaks as a function of operating [...] Read more.
The evolution of the mechanical properties of Zr-2.5Nb pressure tubing during irradiation is dependent on dislocation loop densities that are represented by the broadening of X-ray diffraction lines. Empirical models for the integral breadth of the diffraction peaks as a function of operating conditions have been developed to predict the mechanical properties of CANDU reactor pressure tubes as a function of fast neutron flux, time and temperature. Apart from predicting mechanical property changes based on integral breadth measurements, a new model has been developed to retrospectively deduce abnormal operating temperatures of ex-service pressure from the measured line broadening. The application of integral breadth measurements to assess mechanical properties and temperature variations in pressure tubes is described and discussed in terms of the implications for pressure tube integrity. Full article
(This article belongs to the Special Issue Manufacture, Properties and Applications of Advanced Nuclear Alloys)
Show Figures

Figure 1

Figure 1
<p>Evolution of &lt;a&gt;-type dislocation structure (<b>a</b>,<b>b</b>) and c-component dislocation structure (<b>c</b>,<b>d</b>) in Zr-2.5Nb pressure tubing after irradiation to a fluence of 1.1 × 10<sup>26</sup> n.m<sup>−2</sup> at 270 °C.</p>
Full article ">Figure 2
<p>Mechanical properties of Zr-2.5Nb pressure tubing correlated with radiation damage density from the measurement of prism diffraction line integral breadths (IBs): (<b>a</b>) ultimate tensile strength (UTS); (<b>b</b>) fracture toughness (FT).</p>
Full article ">Figure 3
<p>Schematic showing (<b>a</b>) sample preparation and (<b>b</b>) idealised texture of Zr-2.5Nb pressure tubing and orientation of XRD specimens relative to tube axes.</p>
Full article ">Figure 4
<p>Second-order <math display="inline"><semantics> <mrow> <mfenced open="{" close="}" separators="|"> <mrow> <mn>20</mn> <mover accent="true"> <mrow> <mn>2</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </mfenced> </mrow> </semantics></math> XRD line profiles for the Prism 1 planes of a single crystal compared to a cold-worked Zr-2.5Nb pressure tubing.</p>
Full article ">Figure 5
<p>Integral breadth (IB) of a reconstituted diffraction peak in reciprocal space for the <math display="inline"><semantics> <mrow> <mfenced open="{" close="}" separators="|"> <mrow> <mn>30</mn> <mover accent="true"> <mrow> <mn>3</mn> </mrow> <mo>¯</mo> </mover> <mn>0</mn> </mrow> </mfenced> </mrow> </semantics></math> diffraction line of irradiated single-crystal Zr after deconvolution against the unirradiated profile.</p>
Full article ">Figure 6
<p>Prism plane line broadening (integral breadth) measured from Zr-2.5Nb pressure tube samples irradiated in OSIRISand the NRU reactor. There is less variability in the OSIRIS data because they come from standard fracture toughness specimens that were taken from the same location of a pressure tube and thus subject to less spatial variability than normally exists.</p>
Full article ">Figure 7
<p>Variables used for multiple linear regression models to predict either line broadening or operating temperature: (<b>a</b>) full dataset with 106 entries; (<b>b</b>) reduced dataset with 86 entries for fluences &gt; 0.5 × 10<sup>25</sup> n.m<sup>−2</sup> (E &gt; 1 MeV) that avoids the primary transient in microstructure evolution at low doses.</p>
Full article ">Figure 8
<p>Measured line broadening (points) as a function of axial and azimuthal positions for a pressure tube in (<b>a</b>) a 900 MW and (<b>b</b>) a 600 MW reactor. Predicted integral breadths as calculated from Equation (1) (dotted lines) and Equation (2) (solid lines).</p>
Full article ">Figure 9
<p>Predicted operating temperatures (points) given line broadening measurements as a function of axial and azimuthal positions for a pressure tube in (<b>a</b>) a 900 MW and (<b>b</b>) a 600 MW reactor. The linear regression was obtained using a reduced dataset with 86 entries for fluences &gt; 0.5 × 10<sup>25</sup> n.m<sup>−2</sup> (E &gt; 1 MeV), Equation (2). The nominal temperature of the coolant is shown by the solid line.</p>
Full article ">Figure 10
<p>Measured line broadening (points) as a function of axial and azimuthal positions for a pressure tube in (<b>a</b>) a 900 MW and (<b>b</b>) a 600 MW reactor. Predicted line broadening (solid line) from linear regression using full dataset with 106 entries calculated from Equation (4).</p>
Full article ">Figure 11
<p>Predicted line broadening as a function of fast neutron fluence at 250 °C and 300 °C using Equation (4).</p>
Full article ">Figure 12
<p>Measured line broadening (points) as a function of axial and azimuthal (12 and 6 o’clock) positions and at 6 o’clock near the rolled joint burnish marks for a pressure tube in a 600 MW reactor. Predicted line broadening (dotted line) from linear regression using a full dataset with 106 entries calculated from Equation (3). The interpolated integral breadths for unirradiated material are shown by the horizontal dashed lines. Adapted from [<a href="#B42-metals-14-01446" class="html-bibr">42</a>] with permission from Elsevier.</p>
Full article ">Figure 13
<p>Schematic showing a CANDU reactor rolled joint. Approximate distances are shown for the 600 MWe case. A power-law fit shows the nominal fast fluence profile (in units of 10<sup>25</sup> n.m<sup>−2</sup>) at the inlet and outlet. The fluence profile is shifted inboard at the inlet. A nominal transition to pseudo-steady-state loop evolution at 0.4 × 10<sup>25</sup> n.m<sup>−2</sup> is indicated, below which the line broadening is expected to be rapidly changing due to fluence alone. Adapted from [<a href="#B42-metals-14-01446" class="html-bibr">42</a>] with permission from Elsevier.</p>
Full article ">Figure 14
<p>Variation in transverse yield stress at three different test temperatures for Zr-2.5Nb specimens irradiated at different temperatures. The irradiation dose for all samples was &gt;1 × 10<sup>25</sup> n.m<sup>−2</sup> (E &gt; 1 MeV), deemed to be the dose where the radiation damage had mostly saturated.</p>
Full article ">Figure 15
<p>Fracture strength of hydrides and the transverse yield strength of unirradiated Zr-2.5Nb pressure tubing as a function of test temperature. Also shown are the yield strengths for the main body of Zr-2.5Nb pressure tubing irradiated to fluences &gt; 1 × 10<sup>25</sup> n.m<sup>−2</sup> (E &gt; 1 MeV) and the yield strength estimated for the burnish mark region (see text).</p>
Full article ">Figure 16
<p>Schematic of a single CANDU fuel channel (out of 380 in the 600 MWe design) and typical profiles for temperature, neutron flux, diametral creep and XRD line broadening (dislocation density) as a function of axial location.</p>
Full article ">
11 pages, 511 KiB  
Article
Exploring Structural Changes in Ge-Te Amorphous Films Through Small-Angle Neutron Scattering
by Andrea A. Piarristeguy, Raphaël Escalier, Annie Pradel, Viviana Cristiglio and Gabriel J. Cuello
Appl. Sci. 2024, 14(24), 11713; https://doi.org/10.3390/app142411713 - 16 Dec 2024
Viewed by 395
Abstract
The structure of the glassy GexTe1−x system, with x = 0.17, 0.21, 0.28, 0.30, and 0.45, is studied using the small-angle neutron scattering (SANS) technique. The very-low-momentum-transfer region of the diffractogram exhibits distinct behaviour depending on the germanium content. [...] Read more.
The structure of the glassy GexTe1−x system, with x = 0.17, 0.21, 0.28, 0.30, and 0.45, is studied using the small-angle neutron scattering (SANS) technique. The very-low-momentum-transfer region of the diffractogram exhibits distinct behaviour depending on the germanium content. A similar conclusion is drawn from the analysis of the first diffraction peaks observed at higher angles. This system exhibits three composition regions with distinct behaviours: a first zone of low Ge content (up to about 20–25 at.%), a third zone richer in Ge (from about 30 at.% and above), and a second transitional zone between them. These changes are reflected in the parameters that govern Porod’s region, as well as in the region where the first diffraction peaks appear, corroborating previous observations made using other experimental and simulation techniques. Our study provides experimental evidence that could open up new possibilities for conducting simulations using neutron data. The results presented here show that increasing Ge content leads to a strengthening of the intermediate-range order at the expense of a weakening of the short-range order. Full article
Show Figures

Figure 1

Figure 1
<p>SANS diffractograms as determined on the D16 instrument for the five <math display="inline"><semantics> <mrow> <msub> <mi>Ge</mi> <mi>x</mi> </msub> <msub> <mi>Te</mi> <mrow> <mn>1</mn> <mo>−</mo> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> samples, with <span class="html-italic">x</span> = 0.17, 0.21, 0.28, 0.30, and 0.45.</p>
Full article ">Figure 2
<p>The fit of the model given by Equation (<a href="#FD1-applsci-14-11713" class="html-disp-formula">1</a>) to the <math display="inline"><semantics> <msub> <mi>Ge</mi> <mrow> <mn>0.17</mn> </mrow> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi>Te</mi> <mrow> <mn>0.83</mn> </mrow> </msub> </semantics></math> sample data. The fitted parameters and their optimised values are shown in <a href="#applsci-14-11713-t001" class="html-table">Table 1</a> and <a href="#applsci-14-11713-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>The power-law exponent <math display="inline"><semantics> <mi>β</mi> </semantics></math>, as obtained from the fit of the low-<span class="html-italic">Q</span> region, is plotted as a function of Ge content <span class="html-italic">x</span>. The dashed line is a smoothed spline that only serves as a guide for the eye. The red-, white-, and blue-shaded areas correspond to the different zones observed as a function of the Ge content. Labels and arrows indicate the expected values of <math display="inline"><semantics> <mi>β</mi> </semantics></math> for typical structures or behaviours.</p>
Full article ">Figure 4
<p>Positions of the FSDP (left scale) and PP (right scale) as a function of the Ge content <span class="html-italic">x</span>. Dashed lines are smoothed splines that only serve as a guide for the eye. The red-, white-, and blue-shaded areas correspond to the different zones observed as a function of the Ge content.</p>
Full article ">Figure 5
<p>Correlation distances corresponding to the FSDP and PP as a function of Ge content <span class="html-italic">x</span>. Dashed lines are smoothed splines that only serve as a guide for the eye. The red-, white-, and blue-shaded areas correspond to the different zones observed as a function of the Ge content.</p>
Full article ">
11 pages, 4514 KiB  
Article
Phase Evolution of Li-Rich Layered Li-Mn-Ni-(Al)-O Cathode Materials upon Heat Treatments in Air
by Jekabs Grins, Aleksander Jaworski, Leif Olav Jøsang, Jordi Jacas Biendicho and Gunnar Svensson
Materials 2024, 17(24), 6056; https://doi.org/10.3390/ma17246056 - 11 Dec 2024
Viewed by 353
Abstract
The phase evolution of Li-rich Li-Mn-Ni-(Al)-O cathode materials upon heat treatments in the air at 900 °C was studied by X-ray and neutron powder diffraction. In addition, the structures of Li1.26Mn0.61−xAlx Ni0.15O2, x = [...] Read more.
The phase evolution of Li-rich Li-Mn-Ni-(Al)-O cathode materials upon heat treatments in the air at 900 °C was studied by X-ray and neutron powder diffraction. In addition, the structures of Li1.26Mn0.61−xAlx Ni0.15O2, x = 0.0, 0.05, and 0.10, were refined from neutron powder diffraction data. For two-phase mixtures containing a monoclinic Li2MnO3 type phase M and a rhombohedral LiMn0.5Ni0.5O2 type phase R, the structures, compositions, and phase fractions change with heat treatment time. This is realized by the substitution mechanism 3Ni2+ ↔ 2Li+ + 1Mn4+, which enables cation transport between the phases. A whole-powder pattern fitting analysis of size and strain broadening shows that strain broadening dominates. The X-ray domain size increases with heat treatment time and is larger than the sizes of the domains of M and R observed by electron microscopy. For heat-treated samples, the domain size is smaller for R than for M and decreases with increasing Al doping. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>Octahedral illustrations of the monoclinic M phase (<b>left</b>) and rhombohedral R phase (<b>right</b>). The multiplicity and Wyckoff letter for the sites are indicated.</p>
Full article ">Figure 2
<p>Secondary electron images of Ni35-Al01: (<b>a</b>) calcined at 900 °C for 6 h and (<b>b</b>) after milling.</p>
Full article ">Figure 3
<p>Secondary electron images of Al-doped Ni35 samples heat-treated at 900 °C for 3 weeks: (<b>a</b>) Ni35-Al05; (<b>b</b>) Ni35-Al10.</p>
Full article ">Figure 4
<p>(<b>a</b>) The powder pattern of pyrolyzed, but not calcined, Ni35-Al01. The black line corresponds to a fit of an R phase. The sharp peaks and lower reflection markers are for the Si internal standard. (<b>b</b>) The powder patterns for samples at different holding times t and temperatures. The positions of isolated M phase reflections and Si are indicated.</p>
Full article ">Figure 5
<p>Normalized unit cell volumes V of the M (blue) and R (red) phases for different holding times t at 900 °C for Ni35-Al01. The mean volume (black) is calculated using a constant value of 76 mol% M derived from NPD data [<a href="#B2-materials-17-06056" class="html-bibr">2</a>]. The dashed lines correspond to the unit cell parameters found previously for Ni35-Al01 for a calcination time of 6 h [<a href="#B2-materials-17-06056" class="html-bibr">2</a>]. The unit cell volume for the pyrolyzed starting material is shown by the filled green circle.</p>
Full article ">Figure 6
<p>X-ray domain size (<b>left</b>) in Å and strain (<b>right</b>) for M (blue) and R (red) at different heat treatment times t.</p>
Full article ">Figure 7
<p>Normalized unit cell volumes for Al-doped Ni35 samples; M = blue lines; R = red lines; dashed lines = reference [<a href="#B2-materials-17-06056" class="html-bibr">2</a>]; solid lines = after 11 weeks at 900 °C. Unit cell volumes for Al-doped Ni15 samples are shown by filled green circles.</p>
Full article ">Figure 8
<p>X-ray domain size for M (<b>a</b>) and R (<b>b</b>) and strain for M (<b>c</b>) and R (<b>d</b>). No points are shown (i) for domain size when the peak width is instrumental resolution-limited, (ii) for domain size or strain where the M and R phases are not distinguishable in the powder patterns (Al10 for t ≤ 1 week), and (iii) for samples with anisotropic size broadening for R (Al00, Al01, and Al03 for t = 11 weeks). The arrows show points for which no size broadening is found upon subsequent heat treatments.</p>
Full article ">Figure 9
<p>X Observed (red), calculated (black), and difference (blue) NPD patterns for the Rietveld refinement for Ni15-Al10 from NPD data. The lower reflection markers are for Li<sub>2</sub>CO<sub>3</sub>.</p>
Full article ">
12 pages, 836 KiB  
Article
Comparison of Graphites Intercalated with Fluorine as Slow Neutron Reflectors
by Batiste Clavier, Valentin Czamler, Marc Dubois, Killian Henry, Valery Nesvizhevsky and Elodie Petit
Materials 2024, 17(23), 5972; https://doi.org/10.3390/ma17235972 - 6 Dec 2024
Viewed by 421
Abstract
The use of neutron reflectors is an effective method for improving the quality of neutron sources and neutron delivery systems. In this work, we further develop the method based on the Bragg scattering of neutrons in crystals with large interplanar distances. We compare [...] Read more.
The use of neutron reflectors is an effective method for improving the quality of neutron sources and neutron delivery systems. In this work, we further develop the method based on the Bragg scattering of neutrons in crystals with large interplanar distances. We compare samples of differently prepared fluorine intercalated graphites by measuring the total cross section for the interaction of neutrons with the samples, depending on the neutron wavelength. The Brag scattering cross section is expected to be the dominant part of the total cross section in all the cases. The results show that all samples provide high reflection efficiency over the entire range of the so-called “neutron reflectivity gap” and beyond it, and that they also allow for the choosing of the optimal intercalation methods. Full article
Show Figures

Figure 1

Figure 1
<p>Infrared spectra (ATR mode) of the ARC series (<b>a</b>) and of the other graphite fluoride samples (<b>b</b>); <sup>13</sup>C MAS NMR spectra of selected samples (with the spinning rate of 10 KHz) (<b>c</b>).</p>
Full article ">Figure 2
<p>Outline of the experimental setup of the PF1B instrument. The cold neutron beam enters from the ballistic <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> supermirror ballistic guide H113 on the left and passes through the collimation system containing a series of decreasing apertures in a vacuum. When entering the experimental zone, the beam passes through a disc chopper followed by a monitor and the sample. An <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> supermirror behind the sample provides a cut-off wavelength of about 5<span class="html-italic">Å</span> and allows us to significantly reduce the background in the thermal neutron range due to a non-negligible transmission through the chopper disk. The detector was a <sup>3</sup>He-counter with almost 100% efficiency. A translation stage allows us to scan through up to 8 samples per run.</p>
Full article ">Figure 3
<p>Neutron spectrum measured without a sample. Intensity on the vertical axis is in arbitrary units. Neutron wavelength on the horizontal axis is in Å.</p>
Full article ">Figure 4
<p>Total cross sections of interactions of neutrons (<math display="inline"><semantics> <mi>σ</mi> </semantics></math>) with different intercated graphite samples, per atom in arbitrary units. The normalization per atom uses the mass density and chemical composition presented in <a href="#materials-17-05972-t001" class="html-table">Table 1</a>. The results are presented in two figures to improve visibility.</p>
Full article ">Figure 4 Cont.
<p>Total cross sections of interactions of neutrons (<math display="inline"><semantics> <mi>σ</mi> </semantics></math>) with different intercated graphite samples, per atom in arbitrary units. The normalization per atom uses the mass density and chemical composition presented in <a href="#materials-17-05972-t001" class="html-table">Table 1</a>. The results are presented in two figures to improve visibility.</p>
Full article ">
14 pages, 3419 KiB  
Article
Multievent Correlation with Neutron Volume Detectors
by Noah Nachtigall, Andreas Houben and Richard Dronskowski
Quantum Beam Sci. 2024, 8(4), 30; https://doi.org/10.3390/qubs8040030 - 28 Nov 2024
Viewed by 511
Abstract
The development of advanced volume detectors for neutron time-of-flight diffractometers offers exciting new possibilities. This work takes advantage of these advances by implementing a novel data preprocessing algorithm, exemplified for the first time with data acquired during the operation of a singular mounting [...] Read more.
The development of advanced volume detectors for neutron time-of-flight diffractometers offers exciting new possibilities. This work takes advantage of these advances by implementing a novel data preprocessing algorithm, exemplified for the first time with data acquired during the operation of a singular mounting unit of the POWTEX detector placed at the POWGEN instrument (SNS, ORNL, Oak Ridge, TN, USA). Our approach exploits the additional depth information provided by the volume detector needed to correlate multiple neutron events to neutron trajectories of similar origin and probability. By comparing the properties of these trajectories with the expected physical behavior, one may first identify, then label, and ultimately remove unwanted events due to phenomena such as secondary scattering within the sample environment. This capability has the potential to significantly improve the quality and information content of data collected with neutron diffractometers. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch of the single POWTEX detector mounting unit used in the POWTEX@POWGEN experiment, depicting 3D view, top view, and side view. Cartesian coordinates are also included, as well as the unique POWTEX detector coordinates dubbed <span class="html-italic">S</span>, <span class="html-italic">a</span>, <span class="html-italic">e</span>, and <span class="html-italic">N</span>. Since the detector coordinates are multiples of physical, detector-internal building blocks (cathode stripes, anode wires, modules, etc.), they are unitless in principle but geometrically linked to Cartesian or spherical coordinate systems. The important depth direction is <span class="html-italic">a</span> whereas the direction corresponding to 2<span class="html-italic">θ</span> is <span class="html-italic">S</span>. <span class="html-italic">N</span> divides each mounting unit into eight modules, and each module is internally divided into a lower and an upper side named <span class="html-italic">e</span>. This direction corresponds to the texture angle <span class="html-italic">φ</span>. The direction of the primary neutron beam from the source toward the sample is aligned with the positive <span class="html-italic">z</span> direction.</p>
Full article ">Figure 2
<p>(<b>a</b>) Intensity vs. <span class="html-italic">d</span> diffractogram of a diamond powder measurement from the POWTEX@POWGEN experiment (black) with added artificial secondary scatterer intensity (red-shaded area). (<b>b</b>) Top view demonstration of how this secondary scatterer might have been positioned relative to the sample (exaggerated dimensions).</p>
Full article ">Figure 3
<p>(<b>a</b>) Perspective view of a Debye–Scherrer cone as a ring on a conventional flat-plate detector. (<b>b</b>) Orthographic view of a Debye–Scherrer cone, illustrating the multiple pass-throughs in a volume detector (right) versus the single pass-through for a plate detector (center).</p>
Full article ">Figure 4
<p>(<b>a</b>) Top view of the detector (<span class="html-italic">a</span> vs. <span class="html-italic">S</span>) with different exemplary trajectories of equivalently scattered neutrons at different 2<span class="html-italic">θ</span> values (for one arbitrary wavelength). (<b>b</b>) Side view (<span class="html-italic">N</span>/<span class="html-italic">e</span> vs. <span class="html-italic">a</span>) showing a single trajectory’s flight path and how this translates into a depth–intensity expectation (<span class="html-italic">I</span> vs. <span class="html-italic">a</span>). Note that in this simplified picture, in some of the voxels the trajectory corresponding to the black arrow causes intensity, although no interaction with the boron conversion layer (thin lines) occurs, simply because the single arrow does not reflect the angular uncertainty.</p>
Full article ">Figure 5
<p>The color-scaled intensity plotted against the instrument’s voxel identifiers <span class="html-italic">a</span> and <span class="html-italic">S</span> illustrates how reflections “move” along the detector with λ. The two plots show the same reflections around <span class="html-italic">d</span> = 0.47 Å ± 0.05 Å but for two different wavelengths. In addition, the detector design implies that a reflection shift toward higher <span class="html-italic">S</span>-values can be observed at higher <span class="html-italic">a</span>-values.</p>
Full article ">Figure 6
<p>The color-scaled intensity plot against the <span class="html-italic">a</span> and <span class="html-italic">S</span>′ reveals the presence (<b>a</b>) of three straight vertical reflections with a width of two to four stripes but (<b>b</b>) one additional, extensively broadened reflection. The three straight reflections from the primary scatterer exhibit a distinct and consistent decline in intensity. In contrast, the broad contribution from the secondary scatterer is notable with clearly different nature including a markedly different intensity reduction with depth for each <span class="html-italic">S</span>′.</p>
Full article ">Figure 7
<p>Intensity plots showing the original unedited diffraction data (black) and the diffraction data with an added secondary scatterer (red). (<b>a</b>) Normalized intensity plotted against the depth identifier <span class="html-italic">a</span> for a select number of <span class="html-italic">S</span>′ values showing an <span class="html-italic">S</span>′ value without secondary intensity (<span class="html-italic">S</span>′ = 28) and <span class="html-italic">S</span>′ values with secondary intensity (<span class="html-italic">S</span>′ = 28–34). (<b>b</b>) Intensity plotted against the wavelength for one single <span class="html-italic">S</span>′ = 39.</p>
Full article ">Figure 8
<p>Color-scaled intensity plots after removing bad event data. (<b>a</b>) Intensity vs. <span class="html-italic">a</span> and <span class="html-italic">S</span>′ (see <a href="#qubs-08-00030-f006" class="html-fig">Figure 6</a> for comparison). (<b>b</b>) Intensity vs. <span class="html-italic">d</span> diffractogram with the optimized intensity in black, the original data with the secondary scatterer intensity in red and the leftover bad intensity in blue.</p>
Full article ">
22 pages, 8766 KiB  
Article
Residual Stress Distribution in Dievar Tool Steel Bars Produced by Conventional Additive Manufacturing and Rotary Swaging Processes
by Josef Izák, Pavel Strunz, Olena Levytska, Gergely Németh, Jan Šaroun, Radim Kocich, Marek Pagáč and Kostyantyn Tuharin
Materials 2024, 17(23), 5706; https://doi.org/10.3390/ma17235706 - 22 Nov 2024
Viewed by 717
Abstract
The impact of manufacturing strategies on the development of residual stresses in Dievar steel is presented. Two fabrication methods were investigated: conventional ingot casting and selective laser melting as an additive manufacturing process. Subsequently, plastic deformation in the form of hot rotary swaging [...] Read more.
The impact of manufacturing strategies on the development of residual stresses in Dievar steel is presented. Two fabrication methods were investigated: conventional ingot casting and selective laser melting as an additive manufacturing process. Subsequently, plastic deformation in the form of hot rotary swaging at 900 °C was applied. Residual stresses were measured using neutron diffraction. Microstructural and phase analysis, precipitate characterization, and hardness measurement—carried out to complement the investigation—showed the microstructure improvement by rotary swaging. The study reveals that the manufacturing method has a significant effect on the distribution of residual stresses in the bars. The results showed that conventional ingot casting resulted in low levels of residual stresses (up to ±200 MPa), with an increase in hardness after rotary swaging from 172 HV1 to 613 HV1. SLM-manufactured bars developed tensile hoop and axial residual stresses in the vicinity of the surface and large compressive axial stresses (−600 MPa) in the core due to rapid cooling. The subsequent thermomechanical treatment via rotary swaging effectively reduced both the surface tensile (to approximately +200 MPa) and the core compressive residual stresses (to −300 MPa). Moreover, it resulted in a predominantly hydrostatic stress character and a reduction in von Mises stresses, offering relatively favorable residual stress characteristics and, therefore, a reduction in the risk of material failure. In addition to the significantly improved stress profile, rotary swaging contributed to a fine grain (3–5 µm instead of 10–15 µm for the conventional sample) and increased the hardness of the SLM samples from 560 HV1 to 606 HV1. These insights confirm the utility of rotary swaging as a post-processing technique that not only reduces residual stresses but also improves the microstructural and mechanical properties of additively manufactured components. Full article
(This article belongs to the Special Issue Structural Phenomena in Metallic Materials for Demanding Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Powder used for SLM workpieces.</p>
Full article ">Figure 2
<p>Workpieces’ measurement by neutron diffraction: (<b>a</b>) Conv, ø 40.8 mm; (<b>b</b>) Conv+RS, ø 18.8 mm; (<b>c</b>) SLM, 12 × 12 mm<sup>2</sup>; (<b>d</b>) SLM, ø 12 mm; (<b>e</b>) SLM + RS, ø 11.1 mm.</p>
Full article ">Figure 3
<p>Scheme of the scanned sample with the elements of the measurement geometry for all three scanned components of the strain: (<b>a</b>) radial = <span class="html-italic">y</span>-component, (<b>b</b>) hoop = <span class="html-italic">x</span>-component, and (<b>c</b>) axial = <span class="html-italic">z</span>-component. The contours inside and at the edge of the color areas indicate the edges of the imaginary cross-sections of the individual five samples used for the investigation (see <a href="#materials-17-05706-t002" class="html-table">Table 2</a> for sample dimensions).</p>
Full article ">Figure 4
<p>Microstructure of (<b>a</b>) Conv sample, mag. 7000×; (<b>b</b>) Conv + RS sample, 8000×; (<b>c</b>) SLM sample, mag. 6000×; (<b>d</b>) SLM + RS sample, 3000×.</p>
Full article ">Figure 5
<p>Proportion of BC(C,T)/FCC (red/blue) phase: (<b>a</b>) Conv sample, grid 200 × 200 µm<sup>2</sup>, step size 0.4 µm; (<b>b</b>) Conv+RS sample, grid 200 × 200 µm<sup>2</sup>, step size 0.1 µm; (<b>c</b>) SLM sample, grid 200 × 200 µm<sup>2</sup>, step size 0.05 µm; (<b>d</b>) SLM + RS sample, grid 200 × 200 µm<sup>2</sup>, step size 0.1 µm.</p>
Full article ">Figure 6
<p>Precipitates of: (<b>a</b>) Conv sample, mag. 1000×; (<b>b</b>) Conv + RS sample, mag. 6000×; (<b>c</b>) SLM + RS sample, mag. 3000×.</p>
Full article ">Figure 7
<p>The examples of the measured and fitted diffraction peaks. (<b>a</b>) Sample Conv, the axial component; note the small 110 ferritic phase peak/background ratio for this 40.8 mm diameter sample and the necessary long counting time. (<b>b</b>) Sample SLM-circular, the hoop component measurement; both residual 111 austenite (<b>left</b>) and 110 martensite (<b>right</b>) peaks are visible.</p>
Full article ">Figure 8
<p>The determined components of the strain after removing the pseudostrain for two of the samples: (<b>a</b>) No. 2 (Conv + RS) and (<b>b</b>) No. 4 (SLM-circular).</p>
Full article ">Figure 9
<p>Three components of the stress (points with the error bars) throughout the sample diameter (the gauge volume center of the weight is plotted on the horizontal scale) for the Conv samples: (<b>a</b>) stress components along the scanned line in the Conv sample (No. 1), (<b>b</b>) stress components along the scanned line in the Conv + RS sample (No. 2). The vertical scale is intentionally selected ranging from −800 MPa to +1000 MPa in order to facilitate comparison with the additively manufactured samples’ results shown in the next figure.</p>
Full article ">Figure 10
<p>Three components of the stress (points with the error bars) throughout the sample diameter for the AM samples: (<b>a</b>,<b>b</b>)—stress components along two perpendicular scanned lines for the SLM-rectangular sample (No. 3), (<b>c</b>,<b>d</b>)—stress components along two perpendicular scanned lines for the SLM-circular sample (No. 4), and (<b>e</b>,<b>f</b>)—stress components along two perpendicular scanned lines for the SLM + RS sample (No. 5).</p>
Full article ">Figure 10 Cont.
<p>Three components of the stress (points with the error bars) throughout the sample diameter for the AM samples: (<b>a</b>,<b>b</b>)—stress components along two perpendicular scanned lines for the SLM-rectangular sample (No. 3), (<b>c</b>,<b>d</b>)—stress components along two perpendicular scanned lines for the SLM-circular sample (No. 4), and (<b>e</b>,<b>f</b>)—stress components along two perpendicular scanned lines for the SLM + RS sample (No. 5).</p>
Full article ">Figure 11
<p>(<b>a</b>,<b>c</b>) Von Mises stress and (<b>b</b>,<b>d</b>) hydrostatic stress calculated from the three components of the residual elastic stress. (<b>a</b>,<b>b</b>) show the stress for the conventionally manufactured bars while (<b>c</b>,<b>d</b>) show the stress for the additively manufactured bars.</p>
Full article ">
17 pages, 5578 KiB  
Article
Material Model Fidelity Comparison for the Efficacy of Predicting Residual Stresses in L-PBF Additively Manufactured IN718 Components
by David P. Failla, Matthew J. Dantin, Chuyen J. Nguyen and Matthew W. Priddy
Metals 2024, 14(11), 1210; https://doi.org/10.3390/met14111210 - 24 Oct 2024
Viewed by 976
Abstract
Internal state variable models are well suited to predict the effects of an evolving microstructure as a result of metal-based additive manufacturing (MBAM) processes in components with complex features. As advanced manufacturing techniques such as MBAM become increasingly employed, accurate methods for predicting [...] Read more.
Internal state variable models are well suited to predict the effects of an evolving microstructure as a result of metal-based additive manufacturing (MBAM) processes in components with complex features. As advanced manufacturing techniques such as MBAM become increasingly employed, accurate methods for predicting residual stresses are critical for insight into component performance. To this end, the evolving microstructural model of inelasticity (EMMI) is suited to modeling these residual stresses due to its ability to capture the evolution of rate- and temperature-dependent material hardening as a result of the rapid thermal cycling present in MBAM processes. The current effort contrasts the efficacy of using EMMI with an elastic–perfectly plastic (EPP) material model to predict the residual stresses for an Inconel 718 component produced via laser powder bed fusion (L-PBF). Both constitutive models are used within a thermo-mechanical finite element framework and are validated by published neutron diffraction measurements to demonstrate the need for higher-fidelity models to predict residual stresses in complex components. Both EPP and EMMI can qualitatively predict the residual stresses trends induced by the L-PBF local raster scanning effects on the component, but the influence of the temperature-dependent yield and lack of plastic strain hardening allowed EPP to perform similar to EMMI away from free surfaces. EMMI offered the most insight at the free surfaces and around critical component features, but this work also highlights EMMI as a process–property-dependent model that needs be calibrated to specimens produced with a similar reference structure for microstructure evolution effects to be accurately predicted. Full article
Show Figures

Figure 1

Figure 1
<p>Image of L-shape part showing the C2 and C3 cross–sections intersecting the hole of interest to compare with experimental measurements.</p>
Full article ">Figure 2
<p>(<b>a</b>) The mesh used for the thermal model was partitioned into three sections for applying different materials models to each: the part, the powder, and the substrate. The elements associated with the powder were removed for mechanical analysis. (<b>b</b>) Mesh element aspect ratio contour to illustrate quality of mesh from the top-down view. Element density was coarsened in the powder and substrate since they were not the area of interest, but were still needed to capture their effects on the part.</p>
Full article ">Figure 3
<p>The mechanical simulation showing von Mises stresses consists of 4 steps: (<b>a</b>) initialization of substrate and initial boundary conditions, (<b>b</b>) predictions of stresses as a result of the layer-wise printing, (<b>c</b>) stress relaxation predictions from cooling, (<b>d</b>) predictions of stress as part achieves equilibrium after being removed from the substrate.</p>
Full article ">Figure 4
<p>Thermomechanical modeling flow showing transition from CAD to mesh and process parameters and from CAD to g-code to the AMPES-generated event series to create inputs for the thermal and mechanical models.</p>
Full article ">Figure 5
<p>(<b>a</b>) Power and (<b>b</b>) theoretical input volumetric energy density plotted in 3D space as estimated by the generated event series with associated machine process parameters from [<a href="#B25-metals-14-01210" class="html-bibr">25</a>].</p>
Full article ">Figure 6
<p>Predicted thermal history dictated by moving heat source at final increment of a layer during printing of (<b>a</b>) top view and (<b>b</b>) front view cross-section.</p>
Full article ">Figure 7
<p>Predicted EPP and EMMI residual stress contours for stresses in (<b>a</b>) XX, (<b>b</b>) YY, and (<b>c</b>) ZZ directions for cross–section C2.</p>
Full article ">Figure 8
<p>The predicted EPP and EMMI residual stresses (dashed lines) contrasted with the experimentally measured (EXP, solid lines) residual stresses [<a href="#B5-metals-14-01210" class="html-bibr">5</a>] in the (<b>a</b>) XX, (<b>b</b>) YY, and (<b>c</b>) ZZ along the line of measurement on cross–section C2.</p>
Full article ">Figure 9
<p>Predicted EPP and EMMI residual stress contours for stress in XX for cross–section C3.</p>
Full article ">Figure 10
<p>Predicted EPP and EMMI residual stress contours for stress in YY for cross–section C3.</p>
Full article ">Figure 11
<p>Predicted EPP and EMMI residual stress contours for stress in ZZ for cross–section C3.</p>
Full article ">Figure 12
<p>The predicted EPP and EMMI residual stresses (dashed lines) contrasted with the experimentally measured (EXP, solid lines) residual stresses [<a href="#B5-metals-14-01210" class="html-bibr">5</a>] in the (<b>a</b>) XX, (<b>b</b>) YY, and (<b>c</b>) ZZ along the line of measurement on cross–section C3.</p>
Full article ">
15 pages, 2765 KiB  
Review
“Seeing Is Believing”: How Neutron Crystallography Informs Enzyme Mechanisms by Visualizing Unique Water Species
by Qun Wan and Brad C. Bennett
Biology 2024, 13(11), 850; https://doi.org/10.3390/biology13110850 - 22 Oct 2024
Viewed by 766
Abstract
Hydrogen is the lightest atom and composes approximately half of the atomic content in macromolecules, yet their location can only be inferred or predicted in most macromolecular structures. This is because hydrogen can rarely be directly observed by the most common structure determination [...] Read more.
Hydrogen is the lightest atom and composes approximately half of the atomic content in macromolecules, yet their location can only be inferred or predicted in most macromolecular structures. This is because hydrogen can rarely be directly observed by the most common structure determination techniques (such as X-ray crystallography and electron cryomicroscopy). However, knowledge of hydrogen atom positions, especially for enzymes, can reveal protonation states of titratable active site residues, hydrogen bonding patterns, and the orientation of water molecules. Though we know they are present, this vital layer of information, which can inform a myriad of biological processes, is frustratingly invisible to us. The good news is that, even at modest resolution, neutron crystallography (NC) can reveal this layer and has emerged this century as a powerful tool to elucidate enzyme catalytic mechanisms. Due to its strong and coherent scattering of neutrons, incorporation of deuterium into the protein crystal amplifies the power of NC. This is especially true when solvation and the specific participation of key water molecules are crucial for catalysis. Neutron data allow the modeling of all three atoms in water molecules and have even revealed previously unobserved and unique species such as hydronium (D3O+) and deuteroxide (OD) ions as well as lone deuterons (D+). Herein, we briefly review why neutrons are ideal probes for identifying catalytically important water molecules and these unique water-like species, limitations in interpretation, and four vignettes of enzyme success stories from disparate research groups. One of these groups was that of Dr. Chris G. Dealwis, who died unexpectedly in 2022. As a memorial appreciation of his scientific career, we will also highlight his interest and contributions to the neutron crystallography field. As both the authors were mentored by Chris, we feel we have a unique perspective on his love of molecular structure and admiration for neutrons as a tool to query those structures. Full article
Show Figures

Figure 1

Figure 1
<p>Neutrons are a powerful probe of macromolecular structure, revealing positions of hydrogens and thus protonation states of ionizable residues in proteins. 2<span class="html-italic">Fo–Fc</span> nuclear density (contoured at +1.0 σ, shown as blue mesh) of the protonated active site residues and folate substrate in DHFR at pH 4.5. (<b>A</b>) a protonated His residue; (<b>B</b>) a protonated Glu residue; (<b>C</b>) the protonated folate substrate bound in the active site. (All from PDB: 7D6G). Protein atoms are represented as sticks, with narrow diameter sticks for the hydrogens. Water molecules are represented as ball-and-stick. Atoms are colored as: green for carbon, red for oxygen, blue for nitrogen, and white for hydrogen. Bonding distances are shown in Å.</p>
Full article ">Figure 2
<p>Neutrons are ideally suited for the identification of unique water species in macromolecular structure, especially catalytically important solvent molecules in enzymes. 2<span class="html-italic">Fo-Fc</span> nuclear density (contoured at +1.0 σ, shown as magenta mesh) of deuterated solvent (D<sub>3</sub>O<sup>+</sup>, D<sub>2</sub>O, DO<sup>−</sup>) and deuteron (D<sup>+</sup>) species in the XI, PcyA, XynII, and DHFR structures. (<b>A</b>) D<sub>3</sub>O<sup>+</sup> in XI (PDB: 3KCJ), (<b>B</b>) D<sup>+</sup> in XI (PDB: 3QZA), (<b>C</b>) DO<sup>−</sup> in XI (PDB: 3CWH), (<b>D</b>) D<sub>3</sub>O<sup>+</sup> in PcyA (PDB: 4QCD), (<b>E</b>) D<sub>2</sub>O in XynII (PDB: 4S2F), and (<b>F</b>) D<sup>+</sup> in DHFR (PDB: 7D6G). Protein atoms are represented as sticks, with narrow diameter sticks for the hydrogens. Water molecules are represented as ball-and-stick. Atoms are colored as: green or yellow for carbon, red for oxygen, blue for nitrogen, white for hydrogen, and white or magenta for deuterium or deuterons. Bonding distances are shown in Å.</p>
Full article ">
13 pages, 7200 KiB  
Article
Hydrogenation Properties of the Ti45Zr38−xYxNi17 (5 ≤ x ≤ 10) and the Ti45−zYzZr38Ni17 (5 ≤ z ≤ 15) Mechanically Alloyed Materials
by Joanna Czub, Akito Takasaki, Andreas Hoser, Manfred Reehuis and Łukasz Gondek
Materials 2024, 17(20), 4946; https://doi.org/10.3390/ma17204946 - 10 Oct 2024
Viewed by 669
Abstract
The amorphous materials of the Ti45Zr38Ni17 composition synthesized by mechanical alloying are widely recognized for their ability to store hydrogen with gravimetric densities above 2 wt.%. It is also known that those alloys can form a quasicrystalline state [...] Read more.
The amorphous materials of the Ti45Zr38Ni17 composition synthesized by mechanical alloying are widely recognized for their ability to store hydrogen with gravimetric densities above 2 wt.%. It is also known that those alloys can form a quasicrystalline state after thermal treatment and their structural and hydrogen sorption properties can be altered by doping with various elements. Therefore, in this paper, the results of the studies on the Ti45Zr38Ni17 system with yttrium substituted for titanium and zirconium are presented. We demonstrate that these alloys are able to absorb hydrogen with a concentration of up to 2.7 wt.% while retaining their amorphous structure and they transform into the unique glassy-quasicrystal phase upon annealing. Furthermore, we demonstrate that the in-situ hydrogenation of those new materials is an effortless procedure in which the decomposition of the alloy can be avoided. Moreover, we prove that, in that process, hydrogen does not bind to any specific component of the alloy, which would otherwise cause the formation of simple hydrides or nanoclusters. Full article
Show Figures

Figure 1

Figure 1
<p>The scanning electron microscopy images for the Ti<sub>45</sub>Zr<sub>38−x</sub>Y<sub>x</sub>Ni<sub>17</sub> amorphous alloys.</p>
Full article ">Figure 2
<p>The maps of the elemental distribution for the Ti<sub>45</sub>Zr<sub>38−x</sub>Y<sub>x</sub>Ni<sub>17</sub> amorphous alloys.</p>
Full article ">Figure 3
<p>The scanning electron microscopy images for the Ti<sub>45−z</sub>Y<sub>z</sub>Zr<sub>38</sub>Ni<sub>17</sub> amorphous alloys.</p>
Full article ">Figure 4
<p>The maps of the elemental distribution for the Ti<sub>45−z</sub>Y<sub>z</sub>Zr<sub>38</sub>Ni<sub>17</sub> amorphous alloys.</p>
Full article ">Figure 5
<p>The X-ray diffraction patterns for the Ti<sub>45</sub>Zr<sub>28</sub>Y<sub>10</sub>Ni<sub>17</sub> alloy and its hydride at the maximum hydrogen concentration.</p>
Full article ">Figure 6
<p>The X-ray diffraction patterns for the hydrides Ti<sub>45</sub>Zr<sub>33</sub>Y<sub>5</sub>Ni<sub>17</sub> and the Ti<sub>45</sub>Zr<sub>28</sub>Y<sub>10</sub>Ni<sub>17</sub> alloys at the maximum hydrogen concentrations.</p>
Full article ">Figure 7
<p>The X-ray diffraction patterns for the Ti<sub>45−z</sub>Y<sub>z</sub>Zr<sub>38</sub>Ni<sub>17</sub> (z = 5, 10, 15) alloys.</p>
Full article ">Figure 8
<p>The kinetics of the hydrogenation reaction for the Ti<sub>45</sub>Zr<sub>33</sub>Y<sub>5</sub>Ni<sub>17</sub> and Ti<sub>45</sub>Zr<sub>28</sub>Y<sub>10</sub>Ni<sub>17</sub> alloys at 40 bar and 120 °C.</p>
Full article ">Figure 9
<p>The neutron diffraction patterns for the Ti<sub>45</sub>Zr<sub>33</sub>Y<sub>5</sub>Ni<sub>17</sub> and the Ti<sub>45</sub>Zr<sub>28</sub>Y<sub>10</sub>Ni<sub>17</sub> amorphous alloys and the corresponding deuterides.</p>
Full article ">Figure 10
<p>The neutron diffraction patterns collected during heating of the Ti<sub>40</sub>Y<sub>5</sub>Zr<sub>38</sub>Ni<sub>17</sub>, Ti<sub>35</sub>Y<sub>10</sub>Zr<sub>38</sub>Ni<sub>17</sub>, and Ti<sub>30</sub>Y<sub>15</sub>Zr<sub>38</sub>Ni<sub>17</sub> deuterides.</p>
Full article ">Figure 11
<p>The in-situ neutron diffraction patterns for the Ti<sub>40</sub>Y<sub>5</sub>Zr<sub>38</sub>Ni<sub>17</sub>, Ti<sub>35</sub>Y<sub>10</sub>Zr<sub>38</sub>Ni<sub>17</sub>, and Ti<sub>30</sub>Y<sub>15</sub>Zr<sub>38</sub>Ni<sub>17</sub> amorphous alloys under deuterium pressure of 50 bar.</p>
Full article ">
25 pages, 3993 KiB  
Article
Structural and Dynamical Effects of the CaO/SrO Substitution in Bioactive Glasses
by Margit Fabian, Matthew Krzystyniak, Atul Khanna and Zsolt Kovacs
Molecules 2024, 29(19), 4720; https://doi.org/10.3390/molecules29194720 - 5 Oct 2024
Viewed by 952
Abstract
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, [...] Read more.
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, the effect of SrO/CaO–CaO/SrO substitution on the local structure and dynamics of Si-Na-P-Ca-O oxide glasses has been studied in this work. Differential thermal analysis has been performed to determine the thermal stability of the glasses after the addition of strontium. The local structure has been studied by neutron diffraction augmented by Reverse Monte Carlo simulation, and the local dynamics by neutron Compton scattering and Raman spectroscopy. Differential thermal analysis has shown that SrO-containing glasses have lower glass transition, melting, and crystallisation temperatures. Moreover, the addition of the Sr2+ ions decreased the thermal stability of the glass structure. The total neutron diffraction augmented by the RMC simulation revealed that Sr played a similar role as Ca in the glass structure when substituted on a molar basis. The bond length and the coordination number distributions of the network modifiers and network formers did not change when SrO (x = 0.125, 0.25) was substituted for CaO (25-x). However, the network connectivity increased in glass with 12.5 mol% CaO due to the increased length of the Si-O-Si interconnected chain. The analysis of Raman spectra revealed that substituting CaO with SrO in the glass structure dramatically enhances the intensity of the high-frequency band of 1110–2000 cm−1. For all glasses under investigation, the changes in the relative intensities of Raman bands and the distributions of the bond lengths and coordination numbers upon the SrO substitution were correlated with the values of the widths of nuclear momentum distributions of Si, Na, P, Ca, O, and Sr. The widths of nuclear momentum distributions were observed to soften compared to the values observed and simulated in their parent metal-oxide crystals. The widths of nuclear momentum distributions, obtained from fitting the experimental data to neutron Compton spectra, were related to the amount of disorder of effective force constants acting on individual atomic species in the glasses. Full article
Show Figures

Figure 1

Figure 1
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 1 Cont.
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2 Cont.
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 3
<p>Experimental and RMC calculated structure factors of Ca25 (red), Ca 12.5(blue), and Ca0 (green) glass samples. Curves are displaced by 1 unit successively for clarity. See text for details.</p>
Full article ">Figure 4
<p>Partial atomic pair correlations for Si-O (<b>a</b>), P-O (<b>b</b>), Ca-O (<b>c</b>), Sr-O (<b>d</b>), Na-O (<b>e</b>), and O-O (<b>f</b>), in glassy samples (Ca25 (red), Ca12.5 (blue), Ca0 (green)). The peak positions corresponding to key bond lengths are shown. See text for details.</p>
Full article ">Figure 5
<p>Si-O (<b>a</b>), P-O (<b>b</b>), CaO (<b>c</b>), and O-O (<b>d</b>) coordination number distributions in the glassy samples.</p>
Full article ">Figure 6
<p>Effects of replacement of CaO with SrO on the Raman spectra of phosphosilicate glasses. See text for details.</p>
Full article ">Figure 7
<p>Fits of the TOF spectra recorded at VESUVIO for bioactive glasses Ca25, Ca12.5, and Ca0. Recoil peaks of individual atomic species in the glasses have been colour-coded, with the recoil peaks due to the aluminium container marked in blue. See text for details.</p>
Full article ">Figure 8
<p>The disorder-induced softening of the widths of nuclear momentum distributions of individual atomic species present in bioactive glasses Ca0, Ca12.5, and Ca25. The bar charts show the adopted disorder scale: (i) the blue bars designate the Maxwell-Boltzmann distribution width limits for completely disordered gas of non-interacting particles without an underlying potential; (ii) the white bars show the upper distribution width limits calculated from atom-projected vibrational densities of states of parent metal oxides; and (iii) the red bars show the widths obtained from the analysis of the NCS experiments. See text for details.</p>
Full article ">
16 pages, 8801 KiB  
Article
Multiscale Approach of Investigating the Density of Simulated Fuel for a Zero Power Reactor
by Suneela Sardar, Claude Degueldre and Sarah Green
J. Nucl. Eng. 2024, 5(3), 420-435; https://doi.org/10.3390/jne5030026 - 20 Sep 2024
Viewed by 1032
Abstract
With growing interest in molten salts as possible nuclear fuel systems, knowledge of thermophysical properties of complex salt mixtures, e.g., NaCl-CeCl3, NaCl-UCl3 and NaCl-UCl4, informs understanding and performance modelling of the zero power salt reactor. Fuel density is [...] Read more.
With growing interest in molten salts as possible nuclear fuel systems, knowledge of thermophysical properties of complex salt mixtures, e.g., NaCl-CeCl3, NaCl-UCl3 and NaCl-UCl4, informs understanding and performance modelling of the zero power salt reactor. Fuel density is a key parameter that is examined in a multiscale approach in this paper. In the zero power reactor ‘core’ (cm level), the relative fuel density is estimated for the fuel pin disposition, as well as a function of their pitch (strong effect). Fuel density of the ‘pellet’ (mm–µm level) is first estimated on a geometrical basis, then through tracking pores and cracks using 2D (SEM) and 3D (laser microscopy, LM) techniques. For the nanoscale level, ‘grains’ analysis is done using X-ray diffraction (XRD), revealing the defects, vacancies and swelled grains. Initially, emphasis is on the near-eutectic composition of salt mixtures of CeCl3 with NaCl as the carrier salt. Cerium trichloride (CeCl3) is an inactive surrogate of UCl3 and PuCl3. The results were measured for the specific salt mixture (70 mol% NaCl and 30 mol% CeCl3) in this work, establishing that microscopy and XRD are important techniques for measurement of the physical properties of salts component pellets. This work is of significance, as densities of fuel components affect the power evolution through reactivity and the average neutronic behaviour in zero power salt reactors. Full article
Show Figures

Figure 1

Figure 1
<p>Phase diagram of the NaCl-CeCl<sub>3</sub> system adapted from Lu et al. [<a href="#B19-jne-05-00026" class="html-bibr">19</a>], copyright 2019, with permission from Elsevier.</p>
Full article ">Figure 2
<p>Cross-section through packing of the fuel pin bunch: (<b>a</b>) hexagonal system; (<b>b</b>) quadratic system.</p>
Full article ">Figure 3
<p>Relative fuel density as a function of half pitch for quadratic (<span class="html-italic">D</span><sub>q</sub>) and hexagonal (<span class="html-italic">D</span><sub>h</sub>) systems: (<b>a</b>) lin-lin; (<b>b</b>) log-log. Radius of the fuel pellet: 3.5 mm. Cladding and coating domains are underlined.</p>
Full article ">Figure 3 Cont.
<p>Relative fuel density as a function of half pitch for quadratic (<span class="html-italic">D</span><sub>q</sub>) and hexagonal (<span class="html-italic">D</span><sub>h</sub>) systems: (<b>a</b>) lin-lin; (<b>b</b>) log-log. Radius of the fuel pellet: 3.5 mm. Cladding and coating domains are underlined.</p>
Full article ">Figure 4
<p>SEM micrographs of a NaCl-CeCl<sub>3</sub> powder pressed pellet mixture of composition of NaCl (70 mol%) and CeCl<sub>3</sub> (30 mol%). (<b>a</b>) Indicative view, backscattered contrast, grey scale. (<b>b</b>) Backscattered contrast of the view (<b>a</b>): NaCl (turquoise), CeCl<sub>3</sub> (pink) and porosity/voids (background, dark blue). (<b>c</b>) Indicative grey-scale grey image (higher magnification). (<b>d</b>) Backscattered contrast: NaCl (turquoise), CeCl<sub>3</sub> (pink) and porosity/voids (background, dark blue).</p>
Full article ">Figure 5
<p>Normalised pore size distribution from SEM images (<b>a</b>) and normalised distribution slope (<b>b</b>).</p>
Full article ">Figure 6
<p>SEM-EDX mapping of NaCl-CeCl<sub>3</sub> mixture pellet.</p>
Full article ">Figure 7
<p>Three-dimensional laser microscopy of the NaCl-CeCl<sub>3</sub> mixture pellet: (<b>a</b>) coloured image of the interface, (<b>b</b>) peaks on the surface, (<b>c</b>) valleys, (<b>d</b>) peaks and valleys, (<b>e</b>) height map, (<b>f</b>) height profile showing valleys below the average surface; composition NaCl (70 mol%) and CeCl<sub>3</sub> (30 mol%).</p>
Full article ">Figure 8
<p>Three-dimensional reconstructed images of (<b>a</b>) SEM micrographs of the NaCl-CeCl<sub>3</sub> mixture of <a href="#jne-05-00026-f004" class="html-fig">Figure 4</a>a, (<b>b</b>) SEM micrographs of the NaCl-CeCl<sub>3</sub> mixture of <a href="#jne-05-00026-f004" class="html-fig">Figure 4</a>c, (<b>c</b>) LM micrograph of the NaCl-CeCl<sub>3</sub> mixture of <a href="#jne-05-00026-f007" class="html-fig">Figure 7</a>a and (<b>d</b>) LM micrograph of the NaCl-CeCl<sub>3</sub> mixture of <a href="#jne-05-00026-f007" class="html-fig">Figure 7</a>a.</p>
Full article ">Figure 9
<p>XRD scan of a NaCl-CeCl<sub>3</sub> mixture pellet, along with the peak positions indexed using pellet material references. Composition: NaCl (70 mol%) and CeCl<sub>3</sub> (30 mol%).</p>
Full article ">Figure 10
<p>Peak detail analysis of the scans obtained from the NaCl-CeCl<sub>3</sub> mixture pellet with the pure materials reference patterns. Composition: NaCl (70 mol%) and CeCl<sub>3</sub> (30 mol%).</p>
Full article ">
13 pages, 5698 KiB  
Article
Understanding the High-Temperature Deformation Behaviors in Additively Manufactured Al6061+TiC Composites via In Situ Neutron Diffraction
by Minglei Qu, Dunji Yu, Lianyi Chen, Ke An and Yan Chen
Metals 2024, 14(9), 1064; https://doi.org/10.3390/met14091064 - 17 Sep 2024
Viewed by 1043
Abstract
Aluminum matrix composites (AMCs) are designed to enhance the performance of conventional aluminum alloys for engineering applications at both room and elevated temperatures. However, the dynamic phase-specific deformation behavior and load-sharing mechanisms of AMCs at elevated temperatures have not been extensively studied and [...] Read more.
Aluminum matrix composites (AMCs) are designed to enhance the performance of conventional aluminum alloys for engineering applications at both room and elevated temperatures. However, the dynamic phase-specific deformation behavior and load-sharing mechanisms of AMCs at elevated temperatures have not been extensively studied and remain unclear. Here, in situ neutron diffraction experiments are employed to reveal the phase-specific structure evolution of additively manufactured Al6061+TiC composites under compressive loading at 250 °C. It is found that the addition of a small amount of nano-size TiC significantly alters the deformation behavior and increases the strength at 250 °C in comparison to the as-printed Al6061. Unlike the two-stage behavior observed in Al6061, the Al6061+TiC composites exhibit three stages during compression triggered by changes in the interphase stress states. Further analysis of Bragg peak intensity and broadening reveals that the presence of TiC alters the dislocation activity during deformation at 250 °C by influencing dislocation slip planes and promoting dislocation accumulation. These findings provide direct experimental observations of the phase-specific dynamic process in AMCs under deformation at an elevated temperature. The revealed mechanisms provide insights for the future design and optimization of high-performance AMCs. Full article
Show Figures

Figure 1

Figure 1
<p>In situ neutron diffraction setup. (<b>a</b>) A photo of the experimental setup for the in situ high-temperature compression test (side view). (<b>b</b>) Schematic illustration of the in situ neutron experiment optics at VULCAN (top view). The induction heating coil and the thermocouple are not shown in (<b>b</b>) for simplicity.</p>
Full article ">Figure 2
<p>Compressive stress–strain curve for Al6061 and Al6061+TiC composites at 250 °C.</p>
Full article ">Figure 3
<p>In situ neutron diffraction during compressive loading at 250 °C. (<b>a</b>) Diffraction pattern and Rietveld refinement of Al6061+5%TiC before loading at 250 °C. (<b>b</b>) Diffraction pattern evolution of Al6061+5%TiC.</p>
Full article ">Figure 4
<p>Chemistry changes of Al6061+5%TiC during loading at 250 °C. (<b>a</b>) Two-dimensional contour plot showing the peak intensity evolution of TiC {200} and Mg<sub>2</sub>Si {220}. (<b>b</b>) Evolution of the weight fraction of the TiC and Mg<sub>2</sub>Si phases. (<b>c</b>) Stable chemistry of TiC indicated by the evolution of the modeled Al occupancy at the Ti site.</p>
Full article ">Figure 5
<p>Lattice strain evolution during compressive loading of Al6061 and Al6061+TiC at 250 °C. (<b>a</b>–<b>c</b>) hkl-specific lattice strains determined by SPF. (<b>d</b>–<b>f</b>) average phase-specific lattice strains determined by Rietveld refinement of the full diffraction pattern, which were then compared with the lattice strain of {311} via SPF. The large fitting error bars for the TiC phase in Al6061+2%TiC stem from the smaller TiC fraction and thus the weaker peak intensity.</p>
Full article ">Figure 6
<p>Evolution of “phase stress” (<b>a</b>–<b>c</b>), deviatoric stress (<b>d</b>–<b>f</b>), and “hydrostatic stress” (<b>g</b>–<b>i</b>) during compressive loading. The “phase stresses” were calculated on the basis of the lattice strains obtained from the Rietveld refinement.</p>
Full article ">Figure 7
<p>Evolution of peak intensity for different Al lattice planes during compressive loading. (<b>a</b>–<b>c</b>) Evolution of peak intensity during compression at room temperature. (<b>d</b>–<b>f</b>) Evolution of peak intensity during compression at 250 °C. RT represents room temperature. The peak intensity during loading was normalized by dividing it by the initial peak intensity before loading for each lattice plane.</p>
Full article ">Figure 8
<p>Evolution of peak width for different Al lattice planes during compression of (<b>a</b>) Al6061, (<b>b</b>) Al6061+2%TiC, (<b>c</b>) Al6061+5%TiC at 250 °C. The peak width was calculated by FWHM/d. FWHM is the peak width at the half maximum, and d is the lattice spacing. The peak width during loading was normalized by dividing it by the initial peak width before loading for each lattice plane.</p>
Full article ">
Back to TopTop