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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (623)

Search Parameters:
Keywords = ferromagnetic materials

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 5428 KiB  
Article
C/Ni/N Nanocomposites Based on Hydrolysis Lignin: Synthesis, Study of Structural and Magnetic Properties
by Ihor Bordun, Dariusz Calus, Ewelina Szymczykiewicz, Myroslav Malovanyy, Nazar Nahurskyi, Anatoliy Borysiuk and Yuriy Kulyk
Nanomaterials 2024, 14(23), 1886; https://doi.org/10.3390/nano14231886 - 23 Nov 2024
Viewed by 536
Abstract
A two-step method for the synthesis of C/Ni/N nanocomposites based on hydrolysis lignin from wood chemical processing waste is proposed. These nanocomposites were found to have a well-developed porous structure with a wide pore size distribution. It was shown that doping hydrolysis lignin [...] Read more.
A two-step method for the synthesis of C/Ni/N nanocomposites based on hydrolysis lignin from wood chemical processing waste is proposed. These nanocomposites were found to have a well-developed porous structure with a wide pore size distribution. It was shown that doping hydrolysis lignin with urea-derived nitrogen leads to the appearance of ferromagnetic behavior in the carbon material. When nickel chloride was added during pyrolysis, the magnetic behavior of the C/Ni/N composite was provided by superparamagnetic Ni particles less than 30 nm in size and the magnetism of the carbon matrix. The addition of urea during the synthesis of the nanocomposite further promotes better integration of nickel into the carbon structure. According to the results of magnetic studies, the nickel content in the C/Ni/N nanocomposite was 19 wt.% compared to 15 wt.% in the C/Ni nanocomposite. The synthesized nanocomposite was demonstrated to have no residual magnetization, so its particles do not agglomerate after the external magnetic field is removed. Due to this property and the well-developed porous structure, C/Ni/N composites have the potential to be used as catalysts, active electrode materials for autonomous energy sources, and in environmental technologies as magnetically sensitive adsorbents. Full article
Show Figures

Figure 1

Figure 1
<p>General scheme for the synthesis of nanocomposites C/Ni/N.</p>
Full article ">Figure 2
<p>SEM images and EDX-mapping of the chemical element distribution of C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples.</p>
Full article ">Figure 2 Cont.
<p>SEM images and EDX-mapping of the chemical element distribution of C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples.</p>
Full article ">Figure 3
<p>X-ray diffraction patterns of the synthesized nanocomposites C/Ni/N (<b>a</b>), C/Ni (<b>b</b>), and C/N (<b>c</b>).</p>
Full article ">Figure 4
<p>SAXS spectra of synthesized nanocomposites C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) (points—experimental data, solid line—smoothed curve).</p>
Full article ">Figure 5
<p>SAXS curves plotted in Porod coordinates <span class="html-italic">s</span><sup>4</sup><span class="html-italic">·I(s) = f(s</span><sup>4</sup><span class="html-italic">)</span> for C/Ni (<b>a</b>) and C/N (<b>b</b>) samples.</p>
Full article ">Figure 6
<p>Volume distribution functions of effective pore diameters for C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples.</p>
Full article ">Figure 7
<p>Nitrogen adsorption/desorption isotherms for the C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples.</p>
Full article ">Figure 8
<p>Pore size distribution for C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples calculated by the BJH method.</p>
Full article ">Figure 9
<p>Distribution of pores by size for C/Ni/N (<b>a</b>), C/Ni (<b>b</b>), and C/N (<b>c</b>) samples calculated by the MP method.</p>
Full article ">Figure 10
<p>Remagnetization curve of C/Ni/N nanocomposite. The inset shows a larger scale graph.</p>
Full article ">Figure 11
<p>Magnetic moment hysteresis curve of the C/Ni nanocomposite. The inset shows a larger scale graph.</p>
Full article ">Figure 12
<p>Temperature dependence of the saturation specific magnetization for C/Ni/N (<b>a</b>), C/Ni (<b>b</b>) and C/N (<b>c</b>) samples.</p>
Full article ">Figure 13
<p>Temperature dependences of the saturation specific magnetization: 1 of C/Ni/N composite, 2, 3—model temperature dependences for Ni nanoparticles and nitrogen-containing carbon C/N, respectively.</p>
Full article ">
17 pages, 3344 KiB  
Article
B-Site-Ordered and Disordered Structures in A-Site-Ordered Quadruple Perovskites RMn3Ni2Mn2O12 with R = Nd, Sm, Gd, and Dy
by Alexei A. Belik, Ran Liu, Masahiko Tanaka and Kazunari Yamaura
Molecules 2024, 29(23), 5488; https://doi.org/10.3390/molecules29235488 - 21 Nov 2024
Viewed by 285
Abstract
ABO3 perovskite materials with small cations at the A site, especially with ordered cation arrangements, have attracted a lot of interest because they show unusual physical properties and deviations from general perovskite tendencies. In this work, A-site-ordered quadruple perovskites, RMn3Ni [...] Read more.
ABO3 perovskite materials with small cations at the A site, especially with ordered cation arrangements, have attracted a lot of interest because they show unusual physical properties and deviations from general perovskite tendencies. In this work, A-site-ordered quadruple perovskites, RMn3Ni2Mn2O12 with R = Nd, Sm, Gd, and Dy, were synthesized by a high-pressure, high-temperature method at about 6 GPa. Annealing at about 1500 K produced samples with additional (partial) B-site ordering of Ni2+ and Mn4+ cations, crystallizing in space group Pn–3. Annealing at about 1700 K produced samples with disordering of Ni2+ and Mn4+ cations, crystallizing in space group Im–3. However, magnetic properties were nearly identical for the Pn–3 and Im–3 modifications in comparison with ferromagnetic double perovskites R2NiMnO6, where the degree of Ni2+ and Mn4+ ordering has significant effects on magnetic properties. In RMn3Ni2Mn2O12, one magnetic transition was found at 26 K (for R = Nd), 23 K (for R = Sm), and 22 K (for R = Gd), and two transitions were found at 10 K and 36 K for R = Dy. Curie–Weiss temperatures were close to zero in all compounds, suggesting that antiferromagnetic and ferromagnetic interactions are of the same magnitude. Full article
(This article belongs to the Section Inorganic Chemistry)
Show Figures

Figure 1

Figure 1
<p>Experimental (black crosses), calculated (red line), and difference (blue line at the bottom) room-temperature synchrotron X-ray powder diffraction patterns of NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (in the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K) in a 2<span class="html-italic">θ</span> range of 6° and 59°. The tick marks show possible Bragg reflection positions for the main phase and NiO impurity (from top to bottom). Inset shows a zoomed part in a 2<span class="html-italic">θ</span> range of 16° and 17.8° and emphasizes the presence of the (311) reflection from the B-site ordering. Inset shows a scanning electron microscopy (SEM) image, where the scale bar is 20 µm.</p>
Full article ">Figure 2
<p>Experimental (black crosses), calculated (red line), and difference (blue line at the bottom) room-temperature synchrotron X-ray powder diffraction patterns of NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (in the <span class="html-italic">Im</span>–3 modification, prepared at 1700 K) in a 2<span class="html-italic">θ</span> range of 6° and 59°. The tick marks show possible Bragg reflection positions for the main phase and NiO impurity. Inset shows a zoomed part in a 2<span class="html-italic">θ</span> range of 16° and 17.9° and emphasizes the absence of the (311) reflection and the absence of B-site ordering.</p>
Full article ">Figure 3
<p>(<b>a</b>) The room-temperature cubic lattice parameter in RMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (R = La [<a href="#B28-molecules-29-05488" class="html-bibr">28</a>], Nd, Sm, Gd, and Dy) as a function of the ionic radius R<sup>3+</sup> (for the coordination number 8 as ionic radii for the coordination number XII are not available for small R<sup>3+</sup> cations (R = Gd and Dy) [<a href="#B30-molecules-29-05488" class="html-bibr">30</a>]). NPD: from neutron powder diffraction. XRD: from X-ray powder diffraction. (<b>b</b>) R–O bond length (the left-hand axis) and bond-valence sum for R<sup>3+</sup> (the right-hand axis) in RMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (R = La [<a href="#B28-molecules-29-05488" class="html-bibr">28</a>], Nd, Sm, Gd, and Dy) as a function of the ionic radius R<sup>3+</sup>.</p>
Full article ">Figure 4
<p>(<b>a</b>) ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (<span class="html-italic">χ</span> = <span class="html-italic">M</span>/<span class="html-italic">H</span>) of two modifications of NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K, and the <span class="html-italic">Im</span>–3 modification, prepared at 1700 K) measured at <span class="html-italic">H</span> = 10 kOe. The inset shows the d<span class="html-italic">χT</span>/d<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves (all). (<b>b</b>) ZFC and FCC curves of two modifications of NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> measured at <span class="html-italic">H</span> = 100 Oe. The inset shows the FCC d<span class="html-italic">χT</span>/d<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves.</p>
Full article ">Figure 5
<p>(<b>a</b>) ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (<span class="html-italic">χ</span> = <span class="html-italic">M</span>/<span class="html-italic">H</span>) of two modifications of SmMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K, and the <span class="html-italic">Im</span>–3 modification, prepared at 1700 K) measured at <span class="html-italic">H</span> = 10 kOe. The inset shows FCC d<span class="html-italic">χT</span>/d<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves. (<b>b</b>) ZFC and FCC curves of two modifications of SmMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> measured at <span class="html-italic">H</span> = 100 Oe.</p>
Full article ">Figure 6
<p>ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (<span class="html-italic">χ</span> = <span class="html-italic">M</span>/<span class="html-italic">H</span>) of GdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K) measured at <span class="html-italic">H</span> = 10 kOe. The first inset shows ZFC and FCC curves of GdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> measured at <span class="html-italic">H</span> = 100 Oe. The second inset shows the FCC d<span class="html-italic">χ</span>/d<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves.</p>
Full article ">Figure 7
<p>ZFC (filled symbols) and FCC (empty symbols) dc magnetic susceptibility curves (<span class="html-italic">χ</span> = <span class="html-italic">M</span>/<span class="html-italic">H</span>) of DyMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K) measured at <span class="html-italic">H</span> = 10 kOe (the left-hand axis). The right-hand axis shows the FCC <span class="html-italic">χ</span><sup>−1</sup> versus <span class="html-italic">T</span> curve with the Curie–Weiss fit (black line). The fitting parameters are given in the figure. The inset shows d<span class="html-italic">χ</span>/d<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves.</p>
Full article ">Figure 8
<p>M versus H curves of two modifications of RMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K, and the <span class="html-italic">Im</span>–3 modification, prepared at 1700 K) measured at <span class="html-italic">T</span> = 5 K with (<b>a</b>) R = Nd and (<b>b</b>) R = Sm. The insets show zoomed parts near the origin.</p>
Full article ">Figure 9
<p>M versus H curves of GdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> and DyMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification, prepared at 1500 K) measured at <span class="html-italic">T</span> = 5 K. The inset shows zoomed parts near the origin.</p>
Full article ">Figure 10
<p><span class="html-italic">C</span><sub>p</sub>/<span class="html-italic">T</span> versus <span class="html-italic">T</span> curves of RMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> measured at <span class="html-italic">H</span> = 0 (black curves) and 90 kOe (red curves) for (<b>a</b>) R = Nd (the <span class="html-italic">Pn</span>–3 modification), (<b>b</b>) R = Sm (the <span class="html-italic">Pn</span>–3 modification and the <span class="html-italic">Im</span>–3 modification (blue and brown curves)), (<b>c</b>) R = Gd (the <span class="html-italic">Pn</span>–3 modification), and (<b>d</b>) R = Dy (the <span class="html-italic">Pn</span>–3 modification). Arrows show magnetic transition temperatures. Data below 100 K are shown; inset on panel (<b>a</b>) shows full data up to 270 K (at <span class="html-italic">H</span> = 0 Oe).</p>
Full article ">Figure 11
<p>(<b>a</b>) Real χ′ versus <span class="html-italic">T</span> and (<b>b</b>) imaginary χ″ versus <span class="html-italic">T</span> curves of NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification) at different frequencies (<span class="html-italic">f</span>). Inset shows the χ′ versus <span class="html-italic">T</span> curves at different <span class="html-italic">H</span><sub>ac</sub> (<span class="html-italic">H</span><sub>ac</sub> = 0.05, 0.5, and 5 Oe) and one frequency (<span class="html-italic">f</span> = 300 Hz).</p>
Full article ">Figure 12
<p>Temperature dependence of (<b>a</b>) dielectric constant and (<b>b</b>) loss tangent at different frequencies (<span class="html-italic">f</span>: indicated on the figure) in NdMn<sub>3</sub>Ni<sub>2</sub>Mn<sub>2</sub>O<sub>12</sub> (the <span class="html-italic">Pn</span>–3 modification) at <span class="html-italic">H</span> = 0 Oe. Inset shows frequency dependence of peak positions on loss tangent as <span class="html-italic">T</span><sub>max</sub> versus log(<span class="html-italic">f</span>) (black circles with line) and 1000/<span class="html-italic">T</span><sub>max</sub> versus log(<span class="html-italic">f</span>) (red squares with line).</p>
Full article ">
15 pages, 5423 KiB  
Article
Induction Heating Optimization for Efficient Self-Healing in Asphalt Concrete
by Marina Penalva-Salinas, David Llopis-Castelló, Carlos Alonso-Troyano and Alfredo García
Materials 2024, 17(22), 5602; https://doi.org/10.3390/ma17225602 - 16 Nov 2024
Viewed by 657
Abstract
In this study, the practical application of self-healing asphalt mixtures incorporating steel wool fibers and induction heating was investigated, expanding upon previous research that primarily assessed the self-healing properties rather than optimizing the heating process. Specifically, the aim was to enhance the induction [...] Read more.
In this study, the practical application of self-healing asphalt mixtures incorporating steel wool fibers and induction heating was investigated, expanding upon previous research that primarily assessed the self-healing properties rather than optimizing the heating process. Specifically, the aim was to enhance the induction heating methodology for a semi-dense asphalt concrete mixture (AC 16 Surf 35/50 S). In this research, the induction heating parameters were refined to improve the self-healing capabilities, focusing on the following three key aspects: (i) energy consumption, (ii) heating rate, and (iii) heating homogeneity. The findings reveal that the current intensity, the percentage of ferromagnetic additives, and coil shape are critical for achieving optimal heating conditions. Higher current intensity and additive percentage correlate with improved heating speed and reduced energy consumption. Additionally, variations in coil shape significantly influence the heating uniformity. Although asphalt mixtures with steel slag coarse aggregates exhibit slightly higher specific heat, this aggregate type is preferable for sustainability, as it allows for the recycling of industrial waste. The optimized mixtures can rapidly reach high temperatures, facilitating effective crack repair. This innovation offers a durable, environmentally friendly, and cost-effective solution for road maintenance, thereby enhancing the longevity and performance of asphalt pavements. Full article
(This article belongs to the Special Issue Asphalt Mixtures and Pavements Design (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Particle size distribution.</p>
Full article ">Figure 2
<p>Induction heater.</p>
Full article ">Figure 3
<p>Heating rate characterization: (<b>a</b>) heating process and (<b>b</b>) temperature measurement.</p>
Full article ">Figure 4
<p>Coil shapes used to characterize heating homogeneity: (<b>a</b>) single-turn, (<b>b</b>) double-turn, and (<b>c</b>) two-turn centered.</p>
Full article ">Figure 5
<p>Specific heat values of each sample.</p>
Full article ">Figure 6
<p>Influence of type of coarse aggregate: (<b>a</b>) asphalt mixtures with 2% of steel wool fibers and (<b>b</b>) asphalt mixtures with 4% of steel wool fibers.</p>
Full article ">Figure 7
<p>Influence of the content of steel wool fibers: (<b>a</b>) asphalt mixtures with porphyry coarse aggregate and (<b>b</b>) asphalt mixtures with steel slag coarse aggregate.</p>
Full article ">Figure 8
<p>Homogeneity of heating.</p>
Full article ">Figure 9
<p>Influence of type of coil: (<b>a</b>) single-turn coil, (<b>b</b>) double-turn coil, and (<b>c</b>) two-turn centered coil.</p>
Full article ">
34 pages, 4568 KiB  
Review
Nanothermodynamics: There’s Plenty of Room on the Inside
by Ralph V. Chamberlin and Stuart M. Lindsay
Nanomaterials 2024, 14(22), 1828; https://doi.org/10.3390/nano14221828 - 15 Nov 2024
Viewed by 433
Abstract
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical [...] Read more.
Nanothermodynamics provides the theoretical foundation for understanding stable distributions of statistically independent subsystems inside larger systems. In this review, it is emphasized that extending ideas from nanothermodynamics to simplistic models improves agreement with the measured properties of many materials. Examples include non-classical critical scaling near ferromagnetic transitions, thermal and dynamic behavior near liquid–glass transitions, and the 1/f-like noise in metal films and qubits. A key feature in several models is to allow separate time steps for distinct conservation laws: one type of step conserves energy and the other conserves momentum (e.g., dipole alignment). This “orthogonal dynamics” explains how the relaxation of a single parameter can exhibit multiple responses such as primary, secondary, and microscopic peaks in the dielectric loss of supercooled liquids, and the crossover in thermal fluctuations from Johnson–Nyquist (white) noise at high frequencies to 1/f-like noise at low frequencies. Nanothermodynamics also provides new insight into three basic questions. First, it gives a novel solution to Gibbs’ paradox for the entropy of the semi-classical ideal gas. Second, it yields the stable equilibrium of Ising’s original model for finite-sized chains of interacting binary degrees of freedom (“spins”). Third, it confronts Loschmidt’s paradox for the arrow of time, showing that an intrinsically irreversible step is required for maximum entropy and the second law of thermodynamics, not only in the thermodynamic limit but also in systems as small as N=2 particles. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Finite-size thermal effects. Inset gives Hill’s fundamental equation of small-system thermodynamics, with a simple (three-energy-level) diagram for each term (adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>]). The first three terms on the right side (black) give the standard ways to increase the total internal energy of a system: add heat (<math display="inline"><semantics> <mrow> <mi>T</mi> <mi>d</mi> <msub> <mrow> <mi>S</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), do work on the system (<math display="inline"><semantics> <mrow> <mo>−</mo> <mi>P</mi> <msub> <mrow> <mi>V</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>), or add particles (<math display="inline"><semantics> <mrow> <mi>μ</mi> <mi>d</mi> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>&gt;</mo> <mn>0</mn> </mrow> </semantics></math>). The fourth term (red) contains finite-size effects (surface states, length-scale terms, fluctuations, etc.) that change the width of the levels when the number of subdivisions changes if the subdivision potential is nonzero (<math display="inline"><semantics> <mrow> <mo>ℇ</mo> <mo>≠</mo> <mn>0</mn> </mrow> </semantics></math>). The main figure shows how free energy might change with the number of subdivisions, from <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> in the thermodynamic limit of no subdivisions (<math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>) to <math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi>F</mi> </mrow> <mrow> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">i</mi> <mi mathvariant="normal">n</mi> </mrow> </semantics></math> in the nanothermodynamic limit for stable equilibrium of subsystems inside bulk samples (<math display="inline"><semantics> <mrow> <mi mathvariant="normal">ℇ</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 2
<p>Schematic representation of various multiplicities. A canonical system (<b>top</b>) has two indistinguishable particles that may be on the left side (L), right side (R), or opposite sides. There is only one way to subdivide this system into canonical subsystems (<b>middle</b>), but there are many ways to subdivide it into nanocanonical subsystems (<b>bottom</b>). Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 3
<p>Sketch showing two solutions to Gibbs’ paradox for combining two types of particles: X’s (blue) and O’s (red). (<b>A</b>–<b>C</b>) Canonical ensemble, where all particles of the same type are indistinguishable over all distances. (<b>D</b>–<b>F</b>) Nanocanonical ensemble, comprised of nanoscale subsystems, where similar particles can be distinguished by their location when in different subsystems. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 4
<p>Sketch showing a stable solution of the 1D Ising model at a given <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math>. Ten spins are in the chain. Each spin may be up or down. Each interaction between neighboring spins may be low energy (<math display="inline"><semantics> <mrow> <mo>●</mo> </mrow> </semantics></math>), high energy (<b>X</b>), or a no-energy “break” (<b>O</b>) in the interaction.</p>
Full article ">Figure 5
<p>(<b>D</b>) Temperature dependence of the effective scaling exponent from data (symbols) and models (lines) sketched in (<b>A</b>–<b>C</b>). Each red box encloses a separate set of spins that can be treated using mean-field theory. (<b>A</b>) Standard mean-field theory yields <math display="inline"><semantics> <mrow> <mi>γ</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (dotted line in (<b>D</b>)). (<b>B</b>) Simulations of the standard Ising model yield a monotonic increase in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> with decreasing <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> (dashed line in (<b>D</b>)). (<b>C</b>) The mean-field cluster model yields non-monotonic behavior in <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> (solid lines in (<b>D</b>)), similar to measurements on EuO (circles) and Gd (squares). Difficulty in determining <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math> yields uncertainty as <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>→</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> </mrow> </semantics></math>, but not for <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> <mo>[</mo> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>T</mi> <mo>−</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>)</mo> <mo>/</mo> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>C</mi> </mrow> </msub> <mo>]</mo> </mrow> </mrow> <mo>&gt;</mo> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> where <math display="inline"><semantics> <mrow> <mi>γ</mi> </mrow> </semantics></math> of the standard Ising model shows only gradual and monotonic behavior, unlike the measurements. Adapted from [<a href="#B26-nanomaterials-14-01828" class="html-bibr">26</a>].</p>
Full article ">Figure 6
<p>Log-log plot of frequency-dependent loss from the orthogonal Ising model. The loss is deduced from the power spectral density (PSD) using the fluctuation-dissipation theorem. The frequency is normalized by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> to put the microscopic peak at <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <mi>f</mi> <mo>/</mo> <msub> <mrow> <mi>f</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> <mo>~</mo> <mn>12</mn> </mrow> </semantics></math>. Simulations are made on subsystems of two sizes, each at two temperatures, as given in the legends. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 7
<p>Primary response time of glycerol. Abscissa is inverse temperature, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>/</mo> <mi>T</mi> </mrow> </semantics></math>, where <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math> is the mean-field critical temperature. The ordinate in (<b>A</b>) is <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">log</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, and in (<b>B</b>) it comes from a type of Stickel plot [<a href="#B84-nanomaterials-14-01828" class="html-bibr">84</a>] utilizing finite differences of <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, which removes the prefactor and linearize the VFT2 function. Symbols are from measurements (Stickel [<a href="#B85-nanomaterials-14-01828" class="html-bibr">85</a>]). Various lines are from the VFT2 function Equation (6) (black), VFT function (red), and MYEGA function (blue) [<a href="#B86-nanomaterials-14-01828" class="html-bibr">86</a>]. The inset is a sketch of a simple free-energy diagram, containing two minima separated by a barrier. Primary response in the orthogonal Ising model involves fluctuations in energy that open pathways between the minima. Adapted from [<a href="#B57-nanomaterials-14-01828" class="html-bibr">57</a>].</p>
Full article ">Figure 8
<p>1/<span class="html-italic">f</span>-like noise from maintaining maximum entropy during equilibrium fluctuations. (<b>A</b>–<b>E</b>) Sketch of all distinct configurations of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> spins, arranged in order of decreasing alignment from <math display="inline"><semantics> <mrow> <mi>m</mi> <mo>=</mo> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> (<b>top</b>) to <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> (<b>bottom</b>). The multiplicity for the alignment entropy of the subsystem comes from the number of configurations in each box. (<b>F</b>) Temperature-dependent exponent for noise that varies as a function of frequency, <math display="inline"><semantics> <mrow> <mi>P</mi> <mi>S</mi> <mi>D</mi> <mo>(</mo> <mi>f</mi> <mo>)</mo> <mo>∝</mo> <mn>1</mn> <mo>/</mo> <msup> <mrow> <mi>f</mi> </mrow> <mrow> <mi>α</mi> </mrow> </msup> </mrow> </semantics></math>, with the abscissa normalized by <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mn>1</mn> </mrow> </msub> </mrow> </semantics></math>. Solid symbols (color) are from measurements [<a href="#B88-nanomaterials-14-01828" class="html-bibr">88</a>] of noise in thin films for the metals given in the legend. Open symbols (black) are from simulations of a 3D Ising subsystems having <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>27</mn> </mrow> </semantics></math> spins with dynamics utilizing a local bath to maintain maximum entropy during fluctuations in alignment. Solid line is the best linear fit to the simulations, weighted by the inverse variance of each point. Dashed line is from a random fluctuation model [<a href="#B89-nanomaterials-14-01828" class="html-bibr">89</a>]. Adapted from [<a href="#B60-nanomaterials-14-01828" class="html-bibr">60</a>].</p>
Full article ">Figure 9
<p>Influence of energy on the amplitude of alignment fluctuations via orthogonal dynamics. (<b>A</b>–<b>E</b>) Configurations of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math> interactions, arranged in order of decreasing energy. (<b>F</b>) Simulation of energy (<math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> </mrow> </semantics></math>, red) and magnetization (<math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math>, black) as a function of time for the 1D Ising model containing <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> interactions, with a local bath to maintain maximum entropy. Note how the amplitude of fluctuations in <math display="inline"><semantics> <mrow> <mi>m</mi> </mrow> </semantics></math> tends to be slightly larger when <math display="inline"><semantics> <mrow> <mi>u</mi> <mo>/</mo> <mi>J</mi> <mo>&lt;</mo> <mo>−</mo> <mn>0.3</mn> </mrow> </semantics></math>. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 10
<p>Noise power spectral densities from simulations (lines) and measurements (symbols). Solid lines are from fluctuations in alignment of 1D chains of <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>+</mo> <mn>1</mn> </mrow> </semantics></math> Ising spins using orthogonal dynamics while maintaining maximum entropy. Note that <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math> (blue) is a small enough subsystem to show separate Lorentzians in a 1/<span class="html-italic">f</span>-like spectrum, while <math display="inline"><semantics> <mrow> <mi>b</mi> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math> (red) is large enough to show a crossover from white noise at high frequencies (dotted) to 1/<span class="html-italic">f</span>-like noise at low frequencies with an exponent of <math display="inline"><semantics> <mrow> <mi>α</mi> <mo>=</mo> <mn>0.92</mn> </mrow> </semantics></math> (dashed). Symbols are from measurements of flux noise (solid) and tunnel-coupling noise (open) in a qubit [<a href="#B92-nanomaterials-14-01828" class="html-bibr">92</a>]. Each set of measurements has been shifted in amplitude and frequency to match the simulations. Adapted from [<a href="#B25-nanomaterials-14-01828" class="html-bibr">25</a>].</p>
Full article ">Figure 11
<p>Time dependence of entropies per particle (<b>A</b>–<b>E</b>) and inverse effective temperatures (<b>F</b>). Simulations utilize a Creutz-like model of 1D Ising-like spins coupled to a <span class="html-italic">ke</span> bath of Einstein oscillators. Top three left-side graphs show the time-dependence of <math display="inline"><semantics> <mrow> <mi>S</mi> <mo>/</mo> <mo>(</mo> <mi>N</mi> <mi>k</mi> <mo>)</mo> </mrow> </semantics></math> for the spins (<b>C</b>), <span class="html-italic">ke</span> bath (<b>B</b>), and their sum (<b>A</b>) in a large system, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>2</mn> <mo>×</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mn>6</mn> </mrow> </msup> </mrow> </semantics></math>. Symbols come from first averaging 10,000 sweeps, then averaging three separate simulations of each type, with error bars visible if larger than the symbol size. A simulation with irreversible dynamics (red circles) precedes every simulation with reversible dynamics (black squares). Thus, the total entropy always decreases when the dynamics becomes reversible, as indicated by the orange arrow in (<b>A</b>). Furthermore, when the rate of break-change attempts is reduced to 1/10 the rate of spin-change attempts (middle third of every simulation), reversible simulations have an entropy that depends on the dynamics. Right-side graphs show the total entropies, as in (<b>A</b>) but without time-averaging, over a greatly expanded time scale. Here the differences between reversible (black) and irreversible (red) behavior are clearly visible at the start (<b>D</b>) and end (<b>E</b>) of the simulations. The inset shows corresponding differences in the power-spectral densities of the simulations. Symbols in (<b>F</b>) give the logarithm of the ratio of probabilities of neighboring energy levels in the <span class="html-italic">ke</span> bath, <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi mathvariant="normal">ln</mi> </mrow> <mo>⁡</mo> <mrow> <mo>(</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> <mo>/</mo> <msub> <mrow> <mi>p</mi> </mrow> <mrow> <mi>i</mi> <mo>+</mo> <mn>1</mn> </mrow> </msub> <mo>)</mo> </mrow> </mrow> </mrow> </semantics></math>, with <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> (squares), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> (circles), <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> (up triangles), and <math display="inline"><semantics> <mrow> <mi>i</mi> <mo>=</mo> <mn>3</mn> </mrow> </semantics></math> (down triangles). These values are proportional to the difference in inverse effective temperature of the adjacent levels. A single temperature applies only to irreversible dynamics in the thermodynamic limit (red), not for reversible dynamics in this limit (black) nor for irreversible dynamics of small subsystems, <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>128</mn> </mrow> </semantics></math> (green). Adapted from [<a href="#B16-nanomaterials-14-01828" class="html-bibr">16</a>].</p>
Full article ">Figure 12
<p>Fluctuations in potential energy from MD simulations of Lennard–Jones crystals. Main figure shows normalized <span class="html-italic">pe</span> fluctuations for blocks of <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>32</mn> </mrow> </semantics></math> atoms in a system of <math display="inline"><semantics> <mrow> <mi>N</mi> <mo>=</mo> <mn>442,368</mn> </mrow> </semantics></math> atoms as a function interaction cutoff radius, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>, at three temperatures given in the legend. Note that the data (open symbols) tend to be relatively constant (independent of <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> </mrow> </semantics></math>) when interactions are robustly harmonic, having interaction between nearest-neighbor atoms only, <math display="inline"><semantics> <mrow> <mn>1.12</mn> <mo>≈</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>1</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≤</mo> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>≤</mo> <msup> <mrow> <mn>2</mn> </mrow> <mrow> <mn>4</mn> <mo>/</mo> <mn>6</mn> </mrow> </msup> <mo>≈</mo> <mn>1.59</mn> </mrow> </semantics></math>. Insets show the time dependence of energy autocorrelations in blocks (black squares) and energy correlations between nearest-neighbor blocks (red circles). Simulations are made at <math display="inline"><semantics> <mrow> <mi>k</mi> <mi>T</mi> <mo>/</mo> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>=</mo> <mn>0.0005</mn> </mrow> </semantics></math> for blocks containing a single unit cell of the crystal, <math display="inline"><semantics> <mrow> <mi>n</mi> <mo>=</mo> <mn>4</mn> </mrow> </semantics></math>. The lower inset shows that neighboring blocks are positively correlated when all atoms have robustly harmonic interactions (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>1.5</mn> </mrow> </semantics></math>), while the upper inset (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>r</mi> </mrow> <mrow> <mi>c</mi> </mrow> </msub> <mo>=</mo> <mn>2.0</mn> </mrow> </semantics></math>) shows that neighboring blocks are anticorrelated when interactions include second-neighbor atoms that are anharmonic. Adapted from [<a href="#B15-nanomaterials-14-01828" class="html-bibr">15</a>] with permission from Elsevier.</p>
Full article ">
24 pages, 6021 KiB  
Article
Structural, Optical, Magnetic, and Dielectric Investigations of Pure and Co-Doped La0.67Sr0.33Mn1-x-yZnxCoyO3 Manganites with (0.00 < x + y < 0.20)
by Mansour Mohamed, A. Sedky, Abdullah S. Alshammari, Z. R. Khan, M. Bouzidi and Marzook S. Alshammari
Crystals 2024, 14(11), 981; https://doi.org/10.3390/cryst14110981 - 14 Nov 2024
Viewed by 354
Abstract
Here, we report the structural, optical, magnetic, and dielectric properties of La0.67Sr0.33Mn1-x-yZnxCoyO3 manganite with various x and y values (0.025 < x + y < 0.20). The pure and co-doped samples are [...] Read more.
Here, we report the structural, optical, magnetic, and dielectric properties of La0.67Sr0.33Mn1-x-yZnxCoyO3 manganite with various x and y values (0.025 < x + y < 0.20). The pure and co-doped samples are called S1, S2, S3, S4, and S5, with (x + y) = 0.00, 0.025, 0.05, 0.10, and 0.20, respectively. The XRD confirmed a monoclinic structure for all the samples, such that the unit cell volume and the size of the crystallite and grain were generally decreased by increasing the co-doping content (x + y). The opposite was true for the behaviors of the porosity, the Debye temperature, and the elastic modulus. The energy gap Eg was 3.85 eV for S1, but it decreased to 3.82, 3.75, and 3.65 eV for S2, S5, and S3. Meanwhile, it increased and went to its maximum value of 3.95 eV for S4. The values of the single and dispersion energies (Eo, Ed) were 9.55 and 41.88 eV for S1, but they were decreased by co-doping. The samples exhibited paramagnetic behaviors at 300 K, but they showed ferromagnetic behaviors at 10 K. For both temperatures, the saturated magnetizations (Ms) were increased by increasing the co-doping content and they reached their maximum values of 1.27 and 15.08 (emu/g) for S4. At 300 K, the co-doping changed the magnetic material from hard to soft, but it changed from soft to hard at 10 K. In field cooling (FC), the samples showed diamagnetic regime behavior (M < 0) below 80 K, but this behavior was completely absent for zero field cooling (ZFC). In parallel, co-doping of up to 0.10 (S4) decreased the dielectric constant, AC conductivity, and effective capacitance, whereas the electric modulus, impedance, and bulk resistance were increased. The analysis of the electric modulus showed the presence of relaxation peaks for all the samples. These outcomes show a good correlation between the different properties and indicate that co-doping of up to 0.10 of Zn and Co in place of Mn in La:113 compounds is beneficial for elastic deformation, optoelectronics, Li-batteries, and spintronic devices. Full article
(This article belongs to the Special Issue Crystal Structures and Magnetic Interactions of Magnetic Materials)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 2
<p>SEM micrographs and grain size distribution of the La<sub>0.67</sub>Sr<sub>0.33</sub>M samples.</p>
Full article ">Figure 2 Cont.
<p>SEM micrographs and grain size distribution of the La<sub>0.67</sub>Sr<sub>0.33</sub>M samples.</p>
Full article ">Figure 3
<p>EDS compositional analysis of the La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 3 Cont.
<p>EDS compositional analysis of the La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 4
<p>FTIR spectra of La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 5
<p>(<b>a</b>) Optical absorbance versus wavelength of the La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples. (<b>b</b>) (αhυ)<sup>2</sup> against hυ plots of the La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 6
<p>The linear plots between (n<sup>2</sup> − k<sup>2</sup>) and (λ)<sup>2</sup> for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 7
<p>The linear plots between (n<sup>2</sup> − 1)<sup>−1</sup> and (hυ)<sup>2</sup> for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 8
<p>(<b>a</b>). Magnetization against applied field at 300 K for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples. (<b>b</b>). The dependance of Magnetization on the applied field at 10 K for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>). Magnetic hysteresis loops of La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 10
<p>The zero-field-cooled (ZFC) field-cooled (FC) measurements at an applied magnetic field of 100 Oe for S3 where the blocking temperature T<sub>b</sub> = 60 K.</p>
Full article ">Figure 11
<p>(<b>a</b>). The dependance of a real part of dialectic constant on the frequency (f) for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples. (<b>b</b>). Ac conductivity versus frequency for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>). Real and imaginary parts of electric modulus (M<sup>\</sup>, M<sup>\\</sup>) versus frequency for La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 13
<p>(<b>a</b>,<b>b</b>). Cole-Cole plot of La<sub>0.67</sub>Sr<sub>0.33</sub>Mn<sub>1-x-y</sub>Zn<sub>x</sub>Co<sub>y</sub>O<sub>3</sub> samples.</p>
Full article ">Figure 14
<p>Equivalent circuit of RC circuit for single and two successive semicircles.</p>
Full article ">
39 pages, 8550 KiB  
Review
Enhancement of Magnetic Shielding Based on Low-Noise Materials, Magnetization Control, and Active Compensation: A Review
by Yijin Liu, Jianzhi Yang, Fuzhi Cao, Xu Zhang and Shiqiang Zheng
Materials 2024, 17(22), 5469; https://doi.org/10.3390/ma17225469 - 8 Nov 2024
Viewed by 906
Abstract
Magnetic-shielding technologies play a crucial role in the field of ultra-sensitive physical measurement, medical imaging, quantum sensing, etc. With the increasing demand for the accuracy of magnetic measurement, the performance requirements of magnetic-shielding devices are also higher, such as the extremely weak magnetic [...] Read more.
Magnetic-shielding technologies play a crucial role in the field of ultra-sensitive physical measurement, medical imaging, quantum sensing, etc. With the increasing demand for the accuracy of magnetic measurement, the performance requirements of magnetic-shielding devices are also higher, such as the extremely weak magnetic field, gradient, and low-frequency noise. However, the conventional method to improve the shielding performance by adding layers of materials is restricted by complex construction and inherent materials noise. This paper provides a comprehensive review about the enhancement of magnetic shielding in three aspects, including low-noise materials, magnetization control, and active compensation. The generation theorem and theoretical calculation of materials magnetic noise is summarized first, focusing on the development of spinel ferrites, amorphous, and nanocrystalline. Next, the principles and applications of two magnetization control methods, degaussing and magnetic shaking, are introduced. In the review of the active magnetic compensation system, the forward and inverse design methods of coil and the calculation method of the coupling effect under the ferromagnetic boundary of magnetic shield are explained in detail, and their applications, especially in magnetocardiography (MCG) and magnetoencephalogram (MEG), are also mainly described. In conclusion, the unresolved challenges of different enhancement methods in materials preparation, optimization of practical implementation, and future applications are proposed, which provide comprehensive and instructive references for corresponding research. Full article
Show Figures

Figure 1

Figure 1
<p>Enhancement methods of magnetic shielding based on low-noise materials, magnetization control, and active compensation.</p>
Full article ">Figure 2
<p>Schematic of the permeability measurement using voltammetry.</p>
Full article ">Figure 3
<p>Crystal structures of spinel ferrites: (<b>a</b>) normal spinel, (<b>b</b>) inverse spinel, and (<b>c</b>) orthorhombic, each demonstrating the three crystallographic sites [<a href="#B105-materials-17-05469" class="html-bibr">105</a>].</p>
Full article ">Figure 4
<p>Relationship between effective permeability (<span class="html-italic">μ<sub>e</sub></span>) at 1 kHz and saturation flux density (<span class="html-italic">B<sub>s</sub></span>) for soft magnetic materials [<a href="#B127-materials-17-05469" class="html-bibr">127</a>].</p>
Full article ">Figure 5
<p>Magnetization state in the process of degaussing [<a href="#B147-materials-17-05469" class="html-bibr">147</a>].</p>
Full article ">Figure 6
<p>Simulation results of degaussing-field distribution of (<b>a</b>) I-coils, (<b>b</b>) L-coils, and (<b>c</b>) Z-coils.</p>
Full article ">Figure 7
<p>Simulation of degaussing field of (<b>a</b>) distributed I-coils and (<b>b</b>) distributed L-coils.</p>
Full article ">Figure 8
<p>Simplified distributed-type coils of the magnetic shielding booth: (<b>a</b>) photo of the interior; (<b>b</b>) simulation of degaussing field [<a href="#B158-materials-17-05469" class="html-bibr">158</a>].</p>
Full article ">Figure 9
<p>Configuration of the combined degaussing coils: (<b>a</b>) toroidal coil and (<b>b</b>) solenoidal coil [<a href="#B160-materials-17-05469" class="html-bibr">160</a>].</p>
Full article ">Figure 10
<p>Degaussing waveform with different envelope attenuation functions: (<b>a</b>) linear attenuation, (<b>b</b>) second-order attenuation, and (<b>c</b>) logarithmic attenuation.</p>
Full article ">Figure 11
<p>The end magnetization state of degaussing with different (<b>a</b>) frequencies, (<b>b</b>) initial amplitude, and (<b>c</b>) period number [<a href="#B158-materials-17-05469" class="html-bibr">158</a>].</p>
Full article ">Figure 12
<p>Magnetic properties under magnetic shaking with different frequencies and amplitudes. (<b>a</b>) Magnetization curves. (<b>b</b>) Relative permeabilities [<a href="#B42-materials-17-05469" class="html-bibr">42</a>].</p>
Full article ">Figure 13
<p>Stream functions (<b>i</b>) and coil structure (<b>ii</b>) of biplanar coils designed by TFM: (<b>a</b>) <span class="html-italic">B</span><sub>x</sub> coil and (<b>b</b>) <span class="html-italic">B</span><sub>z</sub> coil [<a href="#B183-materials-17-05469" class="html-bibr">183</a>].</p>
Full article ">Figure 14
<p>Irregular gradient coil designed by boundary element method: (<b>a</b>) d<span class="html-italic">B<sub>x</sub></span>/d<span class="html-italic">x</span> coil; (<b>b</b>) d<span class="html-italic">B<sub>y</sub></span>/d<span class="html-italic">y</span> coil; (<b>c</b>) d<span class="html-italic">B<sub>z</sub></span>/d<span class="html-italic">z</span> coil; (<b>d</b>) gradient coil diagram [<a href="#B190-materials-17-05469" class="html-bibr">190</a>].</p>
Full article ">Figure 15
<p>The change in residual field before and after compensation [<a href="#B147-materials-17-05469" class="html-bibr">147</a>].</p>
Full article ">Figure 16
<p>Relationship between the internal residual field and the current in the external coils [<a href="#B193-materials-17-05469" class="html-bibr">193</a>].</p>
Full article ">Figure 17
<p>Equivalent coil array based on image method [<a href="#B183-materials-17-05469" class="html-bibr">183</a>].</p>
Full article ">Figure 18
<p>Compensation coils system in OPM–MEG: (<b>a</b>) biplanar coils and (<b>b</b>) matrix coils [<a href="#B15-materials-17-05469" class="html-bibr">15</a>,<a href="#B204-materials-17-05469" class="html-bibr">204</a>,<a href="#B205-materials-17-05469" class="html-bibr">205</a>].</p>
Full article ">Figure 19
<p>Cylindrical magnetic shield with single end opening for OPM–MCG [<a href="#B208-materials-17-05469" class="html-bibr">208</a>].</p>
Full article ">
10 pages, 830 KiB  
Article
Coexistence of Superconductivity and Magnetic Ordering in the In–Ag Alloy Under Nanoconfinement
by Marina V. Likholetova, Elena V. Charnaya, Evgenii V. Shevchenko, Yurii A. Kumzerov and Aleksandr V. Fokin
Nanomaterials 2024, 14(22), 1792; https://doi.org/10.3390/nano14221792 - 7 Nov 2024
Viewed by 559
Abstract
The impact of the interface phenomena on the properties of nanostructured materials is the focus of modern physics. We studied the magnetic properties of the nanostructured In–Ag alloy confined within a porous glass. The alloy composition was close to the eutectic point in [...] Read more.
The impact of the interface phenomena on the properties of nanostructured materials is the focus of modern physics. We studied the magnetic properties of the nanostructured In–Ag alloy confined within a porous glass. The alloy composition was close to the eutectic point in the indium-rich range of the phase diagram. Temperature dependences of DC magnetization evidenced two superconducting transitions at 4.05 and 3.38 K. The magnetization isotherms demonstrated the superposition of two hysteresis loops with low and high critical fields below the second transition, a single hysteresis between the transitions and ferromagnetism with weak remanence in the normal state of the alloy. The shape of the loop seen below the second transition, which closes at a low magnetic field, corresponded to the intermediate state of the type-I superconductor. It was ascribed to strongly linked indium segregates. The loop observed below the first transition is referred to as type-II superconductivity. The secondary and tertiary magnetization branches measured at decreasing and increasing fields were shifted relative to each other, revealing the proximity of superconducting and ferromagnetic phases at the nanometer scale. This phenomenon was observed for the first time in the alloy, whose components were not magnetic in bulk. The sign of the shift shows the dominant role of the stray fields of ferromagnetic regions. Ferromagnetism was suggested to emerge at the interface between the In and AgIn2 segregates. Full article
(This article belongs to the Section Nanocomposite Materials)
Show Figures

Figure 1

Figure 1
<p>Pore size distribution in the porous glass.</p>
Full article ">Figure 2
<p>Temperature dependences of susceptibilities measured under the ZFC (black symbols and lines), FCC (blue symbols and lines), and FCW (red symbols and lines) protocols at fields of 10 (<b>a</b>), 50 (<b>b</b>), and 100 (<b>c</b>) Oe.</p>
Full article ">Figure 3
<p>Temperature dependences of susceptibilities measured at 300 and 500 Oe under the ZFC (black symbols and lines), FCC (blue symbols and lines), and FCW (red symbols and lines) protocols.</p>
Full article ">Figure 4
<p>Central parts of isotherms of magnetizations obtained at temperatures 1.8 (<b>a</b>), 3.3 (<b>b</b>), 3.6 (<b>c</b>), and 8 (<b>d</b>) K. The arrows indicate the directions of ramping the field. The red, green, and blue symbols and lines correspond to the virgin, secondary, and tertiary magnetizations, respectively. The insets present the magnetization curves on a larger scale.</p>
Full article ">Figure 5
<p>X-ray pattern of the porous glass/In–Ag alloy nanocomposite.</p>
Full article ">Figure 6
<p>The separated central hysteresis loops for temperatures 1.8 (<b>a</b>) and 3.3 (<b>b</b>) K. The arrows indicate the directions of ramping the field. The red, green, and blue symbols and lines correspond to the virgin, secondary, and tertiary magnetizations, respectively.</p>
Full article ">
25 pages, 5238 KiB  
Article
Numerical Simulation of Electromagnetic Nondestructive Testing Technology for Elasto–Plastic Deformation of Ferromagnetic Materials Based on Magneto–Mechanical Coupling Effect
by Xiangyi Hu, Xiaoqiang Wang, Haichao Cai, Xiaokang Yang, Sanfei Pan, Yafeng Yang, Hao Tan and Jianhua Zhang
Sensors 2024, 24(22), 7103; https://doi.org/10.3390/s24227103 - 5 Nov 2024
Viewed by 512
Abstract
A numerical tool for simulating the detection signals of electromagnetic nondestructive testing technology (ENDT) is of great significance for studying detection mechanisms and improving detection efficiency. However, the quantitative analysis methods for ENDT have not yet been sufficiently studied due to the absence [...] Read more.
A numerical tool for simulating the detection signals of electromagnetic nondestructive testing technology (ENDT) is of great significance for studying detection mechanisms and improving detection efficiency. However, the quantitative analysis methods for ENDT have not yet been sufficiently studied due to the absence of an effective constitutive model. This paper proposed a new magneto–mechanical model that can reflect the dependence of relative permeability on elasto–plastic deformation and proposed a finite element–infinite element coupling method that can replace the traditional finite element truncation boundary. The validity of the finite element–infinite element coupling method is verified by the experimental result of testing electromagnetic analysis methods using TEAM Problem 7. Then, the reliability and accuracy of the proposed model are verified by comparing the simulation results under elasto–plastic deformation with experimental results. This paper also investigates the effect of elasto–plastic deformation on the transient magnetic flux signal, a quantitative hyperbolic tangent model between Bzpp (peak–peak value of the normal component of magnetic flux signal) and elastic stress, and the exponential function relationship between Bzpp and plastic deformation is established. In addition, the difference and mechanism of a magnetic flux signal under elasto–plastic deformations are analyzed. The results reveal that the variation of the transient magnetic flux signal is mainly due to domain wall pinning, which is significantly affected by elasto–plastic deformation. The results of this paper are important for improving the accuracy of quantitative ENDT for elasto–plastic deformation. Full article
(This article belongs to the Section Physical Sensors)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of magnetic field and solid mechanics coupling simulation.</p>
Full article ">Figure 2
<p>The computation model of TEAM Problem 7 and the comparison results. (<b>a</b>) Computation model of TEAM Problem 7, (<b>b</b>) Magnetic flux density <span class="html-italic">B<sub>z</sub></span>.</p>
Full article ">Figure 3
<p>Comparison between the inverse of the relative permeability calculated by different models and the experimental results [<a href="#B28-sensors-24-07103" class="html-bibr">28</a>].</p>
Full article ">Figure 4
<p>Comparison between the theoretical calculation results of the proposed model and the experimental data [<a href="#B32-sensors-24-07103" class="html-bibr">32</a>]. (<b>a</b>) Tangential component H<span class="html-italic">x</span>, (<b>b</b>) normal component H<span class="html-italic">z</span>, (<b>c</b>) the average of H<span class="html-italic">x</span>, (<b>d</b>) the slope and maximum of H<span class="html-italic">z</span>.</p>
Full article ">Figure 5
<p>Comparison between the theoretical results and the experimental data [<a href="#B33-sensors-24-07103" class="html-bibr">33</a>]. (<b>a</b>) The tangential component H<span class="html-italic">x</span>, (<b>b</b>) the normal component H<span class="html-italic">z</span>.</p>
Full article ">Figure 6
<p>The geometry and dimension of the simulation model.</p>
Full article ">Figure 7
<p>Signals obtained from simulation results. (<b>a</b>) Excitation voltage, (<b>b</b>) tangential component B<span class="html-italic">x</span>, (<b>c</b>) normal component B<span class="html-italic">z</span>, (<b>d</b>) tangential component B1<span class="html-italic">x</span>, (<b>e</b>) normal component B1<span class="html-italic">z</span>.</p>
Full article ">Figure 8
<p>B-H curve and nonlinear magnetic signal variation based on the excitation voltage signal. (<b>a</b>) Excitation voltage signal, (<b>b</b>) B-H curve, (<b>c</b>) tangential component, (<b>d</b>) normal component.</p>
Full article ">Figure 9
<p>Effects of elastic stress on magnetic flux density through the coil 2. (<b>a</b>) Tangential component B<span class="html-italic">x</span>, (<b>b</b>) normal component B<span class="html-italic">z</span>, (<b>c</b>) the relationship between B<span class="html-italic">x</span><sub>pp</sub> and elastic stress, (<b>d</b>) the comparison of the relationship between B<span class="html-italic">z</span><sub>pp</sub> or experimental results and elastic stress.</p>
Full article ">Figure 10
<p>Comparison between the theoretical results and the experimental results [<a href="#B36-sensors-24-07103" class="html-bibr">36</a>,<a href="#B37-sensors-24-07103" class="html-bibr">37</a>]. (<b>a</b>) Langman’s experimental results [<a href="#B36-sensors-24-07103" class="html-bibr">36</a>], (<b>b</b>) Kenji’s experimental results [<a href="#B37-sensors-24-07103" class="html-bibr">37</a>].</p>
Full article ">Figure 11
<p>Effects of plastic stress on magnetic flux density through coil 2. (<b>a</b>) Tangential component B<span class="html-italic">x</span>, (<b>b</b>) normal component B<span class="html-italic">z</span>, (<b>c</b>) the relationship between B<span class="html-italic">x</span><sub>pp</sub> and plastic stress, (<b>d</b>) the relationship between B<span class="html-italic">z</span><sub>pp</sub> and plastic stress.</p>
Full article ">Figure 12
<p>Comparison between the quantitative results and simulation or experimental results [<a href="#B40-sensors-24-07103" class="html-bibr">40</a>]. (<b>a</b>) Experimental result of <span class="html-italic">B</span>1<span class="html-italic">z</span><sub>pp</sub> [<a href="#B40-sensors-24-07103" class="html-bibr">40</a>], (<b>b</b>) Experimental result of K factor [<a href="#B40-sensors-24-07103" class="html-bibr">40</a>].</p>
Full article ">Figure A1
<p>Infinite geometry mapping for 3D problems. (<b>a</b>) Sub-element, (<b>b</b>) parent element, the coordinates of point: 1 (−1,1,−1), 2 (−1,−1,−1), 3 (−1,−1,1), 4 (−1,1,1).</p>
Full article ">
17 pages, 3279 KiB  
Article
Theoretical Study of the Magnetic Mechanism of a Pca21 C4N3 Monolayer and the Regulation of Its Magnetism by Gas Adsorption
by Dongqiu Zhao, Xiao Tang, Xueying Gao, Wanyan Xing, Shuli Liu, Huabing Yin and Lin Ju
Molecules 2024, 29(21), 5194; https://doi.org/10.3390/molecules29215194 - 2 Nov 2024
Viewed by 489
Abstract
For metal-free low-dimensional ferromagnetic materials, a hopeful candidate for next-generation spintronic devices, investigating their magnetic mechanisms and exploring effective ways to regulate their magnetic properties are crucial for advancing their applications. Our work systematically investigated the origin of magnetism of a graphitic carbon [...] Read more.
For metal-free low-dimensional ferromagnetic materials, a hopeful candidate for next-generation spintronic devices, investigating their magnetic mechanisms and exploring effective ways to regulate their magnetic properties are crucial for advancing their applications. Our work systematically investigated the origin of magnetism of a graphitic carbon nitride (Pca21 C4N3) monolayer based on the analysis on the partial electronic density of states. The magnetic moment of the Pca21 C4N3 originates from the spin-split of the 2pz orbit from special carbon (C) atoms and 2p orbit from N atoms around the Fermi energy, which was caused by the lone pair electrons in nitrogen (N) atoms. Notably, the magnetic moment of the Pca21 C4N3 monolayer could be effectively adjusted by adsorbing nitric oxide (NO) or oxygen (O2) gas molecules. The single magnetic electron from the adsorbed NO pairs with the unpaired electron in the N atom from the substrate, forming a Nsub-Nad bond, which reduces the system’s magnetic moment from 4.00 μB to 2.99 μB. Moreover, the NO adsorption decreases the both spin-down and spin-up bandgaps, causing an increase in photoelectrical response efficiency. As for the case of O2 physisorption, it greatly enhances the magnetic moment of the Pca21 C4N3 monolayer from 4.00 μB to 6.00 μB through ferromagnetic coupling. This method of gas adsorption for tuning magnetic moments is reversible, simple, and cost-effective. Our findings reveal the magnetic mechanism of Pca21 C4N3 and its tunable magnetic performance realized by chemisorbing or physisorbing magnetic gas molecules, providing crucial theoretical foundations for the development and utilization of low-dimensional magnetic materials. Full article
(This article belongs to the Special Issue Novel Two-Dimensional Energy-Environmental Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The spin-polarized PDOS of C<sub>C</sub>, C<sub>N</sub>, and N 2<span class="html-italic">p</span> in the Pca21 C<sub>4</sub>N<sub>3</sub> monolayer. The C<sub>C</sub> 2<span class="html-italic">p</span>, C<sub>N</sub> 2<span class="html-italic">p</span>, and N 2<span class="html-italic">p</span> states are represented by blue, green, and red lines, respectively. The Fermi level (<span class="html-italic">E</span><sub>f</sub>) is indicated by a black dashed line and set to 0 eV. This representation of <span class="html-italic">E</span><sub>f</sub> is also applicable to the subsequent density of states (DOS) plots. (<b>b</b>) The three-dimensional isosurfaces (iso-value of 0.01 e/Å<sup>3</sup>) depicting net magnetization density (difference between spin-up and spin-down), which also applies to the subsequent net magnetization density, for the Pca21 C<sub>4</sub>N<sub>3</sub> monolayer in the ferromagnetic state. Gray spheres symbolize C atoms, and blue spheres denote N atoms, which also applies to the subsequent Figures 2, 5 and 7. The subfigure labels represent the coordinates.</p>
Full article ">Figure 2
<p>The spin-resolved PDOS diagrams for the 2<span class="html-italic">s</span> and 2<span class="html-italic">p</span> states of one (<b>a</b>) N, (<b>b</b>) C<sub>C</sub>, and (<b>c</b>) C<sub>N</sub> atom in the Pca21 C<sub>4</sub>N<sub>3</sub> monolayer. The 2<span class="html-italic">s</span> and 2<span class="html-italic">p</span> states are represented by blue and red lines, respectively. (<b>d</b>) The ELF of Pca21 C<sub>4</sub>N<sub>3</sub> monolayer, with cyan regions indicating electron accumulation. The isosurface value is set to 0.60 e/Å<sup>3</sup>.</p>
Full article ">Figure 3
<p>The spin-resolved PDOS of the 2<span class="html-italic">p</span><sub>x</sub>, 2<span class="html-italic">p</span><sub>y</sub>, and 2<span class="html-italic">p</span><sub>z</sub> for N (designated as (<b>a</b>) N<sub>17</sub>, (<b>b</b>) N<sub>18</sub>, and (<b>c</b>) N<sub>19</sub>; see <a href="#app1-molecules-29-05194" class="html-app">Figure S1</a>), and (<b>d</b>) C<sub>N</sub>, and (<b>e</b>) C<sub>C</sub> in the Pca21 C<sub>4</sub>N<sub>3</sub> monolayer. For C<sub>C</sub>, C<sub>N</sub>, and N, 2<span class="html-italic">p</span><sub>x</sub>, 2<span class="html-italic">p</span><sub>y</sub> and 2<span class="html-italic">p</span><sub>z</sub> are represented by green, red, and blue lines, respectively.</p>
Full article ">Figure 4
<p>The adsorption energy and net magnetic moment of (<b>a</b>) NO@C<sub>4</sub>N<sub>3</sub> systems and (<b>b</b>) O<sub>2</sub>@C<sub>4</sub>N<sub>3</sub> systems at different adsorption sites. The green lines denote adsorption energy, and the blue lines indicate the values of the magnetic moment.</p>
Full article ">Figure 5
<p>(<b>a</b>) The top (upper) and profile (lower) perspectives of the optimized configuration of the NO@C<sub>4</sub>N<sub>3</sub> system. (<b>b</b>) The spin-resolved PDOS of 2<span class="html-italic">p</span> for (NO)<sub>f</sub>, C<sub>C</sub>, C<sub>N</sub>, and N in the NO@C<sub>4</sub>N<sub>3</sub> system. The 2<span class="html-italic">p</span> of (NO)<sub>f</sub> is denoted by red lines, and the C<sub>C</sub> 2<span class="html-italic">p</span>, C<sub>N</sub> 2<span class="html-italic">p</span>, and N 2<span class="html-italic">p</span> are denoted by blue, green, and pink lines, respectively. (<b>c</b>) In (NO)<sub>i</sub>, the spin-resolved PDOS of N 2<span class="html-italic">p</span> and O 2<span class="html-italic">p</span> is represented by blue and red lines, respectively. The inset displays the spatial distribution of spin-up <span class="html-italic">π</span>* orbitals for (NO)<sub>i</sub>. In (NO)<sub>i</sub>, (<b>d</b>) the PDOS of N 2<span class="html-italic">p</span><sub>x</sub>, 2<span class="html-italic">p</span><sub>y</sub>, and 2<span class="html-italic">p</span><sub>z</sub> is represented by green, rose, and black lines, respectively, and (<b>e</b>) the PDOS of O 2<span class="html-italic">p</span><sub>x</sub>, 2<span class="html-italic">p</span><sub>y</sub>, and 2<span class="html-italic">p</span><sub>z</sub> is represented by orange, purple, and cyan lines, respectively. (<b>f</b>) The views of the 3D isosurfaces (iso-value of 0.01 e/Å<sup>3</sup>) of net magnetization density for the NO@C<sub>4</sub>N<sub>3</sub>. (<b>g</b>) Integrals of CDD along the <span class="html-italic">z</span> direction for the NO@C<sub>4</sub>N<sub>3</sub> system. The inset depicts the CDD distributions, with yellow regions representing electron accumulation and cyan regions indicating electron depletion. The isosurface value is established at 5.00 × 10<sup>−3</sup> e/Å<sup>3</sup>. Red spheres denote O atoms, which also applies to the subsequent Figure 7.</p>
Full article ">Figure 6
<p>The spin-resolved PDOS of N 2<span class="html-italic">p</span> (denoted with red lines) and O 2<span class="html-italic">p</span> (denoted with blue lines) in (<b>a</b>) (NO)<sub>i</sub> and (<b>b</b>) (NO)<sub>f</sub>; (NO)<sub>i</sub> and (NO)<sub>f</sub> denote NO before and after adsorption on Pca21 C<sub>4</sub>N<sub>3</sub>, respectively. The spin-resolved PDOS of 2<span class="html-italic">p</span> for N<sub>sub</sub> (<b>c</b>), C<sub>C-near</sub> (<b>d</b>), and C<sub>N-near</sub> (<b>e</b>) in systems; the red lines labeled with <b>i</b> denote the 2<span class="html-italic">p</span> sates before NO adsorption, and the blue lines labeled with <b>f</b> represent the 2<span class="html-italic">p</span> after NO adsorption. N<sub>sub</sub> refers to the N in the C<sub>4</sub>N<sub>3</sub> substrate bonded with the NO molecule, C<sub>C-near</sub> is C<sub>C</sub> adjacent to N<sub>sub</sub>, and C<sub>N-near</sub> is C<sub>N</sub> neighboring N<sub>sub</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) The PDOSs for (O<sub>2</sub>)<sub>f</sub> and C<sub>4</sub>N<sub>3</sub> in the O<sub>2</sub>@C<sub>4</sub>N<sub>3</sub> system, which are represented by blue and red lines, respectively. (<b>b</b>) Integrals of CDD along the <span class="html-italic">z</span> direction for the O<sub>2</sub>@C<sub>4</sub>N<sub>3</sub> system. The inset depicts the CDD distributions, with yellow regions representing electron accumulation and cyan regions indicating electron depletion. The isosurface value is established at 2.00 × 10<sup>−3</sup> e/Å<sup>3</sup>. (<b>c</b>) The top (upper) and profile (lower) views of the 3D isosurfaces (the iso-value is 1.15 × 10<sup>−2</sup> e/Å<sup>3</sup>) of net magnetization density for the O<sub>2</sub>@C<sub>4</sub>N<sub>3</sub> monolayer. (<b>d</b>) The spin-resolved PDOSs of the 2<span class="html-italic">p</span><sub>x</sub>, 2<span class="html-italic">p</span><sub>y</sub>, and 2<span class="html-italic">p</span><sub>z</sub> for (O<sub>2</sub>)<sub>i</sub>, which are represented by green, red, and blue lines, respectively. The inset displays the spatial distribution of spin-up <span class="html-italic">π</span>* orbitals for (O<sub>2</sub>)<sub>i</sub>.</p>
Full article ">
20 pages, 20556 KiB  
Article
A Contactless Low-Carbon Steel Magnetostrictive Torquemeter: Numerical Analysis and Experimental Validation
by Carmine Stefano Clemente, Claudia Simonelli, Nicolò Gori, Antonino Musolino, Rocco Rizzo, Marco Raugi, Alessandra Torri and Luca Sani
Sensors 2024, 24(21), 6949; https://doi.org/10.3390/s24216949 - 29 Oct 2024
Viewed by 496
Abstract
Torque measurement is a key task in several mechanical and structural engineering applications. Most commercial torquemeters require the shaft to be interrupted to place the sensors between the two portions of the shaft where a torque has to be measured. Contactless torquemeters based [...] Read more.
Torque measurement is a key task in several mechanical and structural engineering applications. Most commercial torquemeters require the shaft to be interrupted to place the sensors between the two portions of the shaft where a torque has to be measured. Contactless torquemeters based on the inverse magnetostrictive effect represent an effective alternative to conventional ones. Most known ferromagnetic materials have an inverse magnetostrictive behavior: applied stresses induce variations in their magnetic properties. This paper investigates the possibility of measuring torsional loads applied to a shaft made of ferromagnetic steel S235 through an inverse magnetostrictive torquemeter. It consists of an excitation coil that produces a time-varying electromagnetic field inside the shaft and an array of sensing coils suitably arranged around it, in which voltages are induced. First, the system is analyzed both in unloaded and loaded conditions by a Finite Element Method, investigating the influence of relative positions between the sensor and the shaft. Then, the numerical results are compared with the experimental measurements, confirming a linear characteristic of the sensor (sensitivity about 0.013 mV/Nm for the adopted experimental setup) and revealing the consistency of the model used. Since the system exploits the physical behavior of a large class of structural steel and does not require the introduction of special materials, this torquemeter may represent a reliable, economical, and easy-to-install device. Full article
(This article belongs to the Special Issue Magnetostrictive Transducers, Sensors, and Actuators)
Show Figures

Figure 1

Figure 1
<p>Schematic view of the torque measurement system.</p>
Full article ">Figure 2
<p>Ampere-turn equivalent model of a sensor with an excitation coil and a single couple of sense coils.</p>
Full article ">Figure 3
<p>Meshed domain for the FE analysis: shaft portion and air box (<b>a</b>), and flux density distribution on the shaft surface (<b>b</b>).</p>
Full article ">Figure 4
<p>Complete sensor over a ferromagnetic shaft in air. The excitation and sensing coils layout over the under-testing shaft is visible. The pickup coils 1 and 4, 2 and 5, and 3 and 6 are electrically connected in series.</p>
Full article ">Figure 5
<p>Amplitudes of the induced voltages as functions of the gap.</p>
Full article ">Figure 6
<p>Amplitudes of the induced voltages with respect to the pitch angle.</p>
Full article ">Figure 7
<p>Amplitudes of the induced voltages with respect to the displacement along the x-axis.</p>
Full article ">Figure 8
<p>Amplitudes of the induced voltages with respect to the yaw angle.</p>
Full article ">Figure 9
<p>FE model of the analyzed geometry.</p>
Full article ">Figure 10
<p>Amplitude of the magnetic flux density in the 3D FE model at 0 Nm (<b>a</b>), 600 Nm (<b>b</b>), and 1200 Nm (<b>c</b>). Values are expressed in milliTesla.</p>
Full article ">Figure 11
<p>FE computed sensing coils peak voltages versus applied pure torque.</p>
Full article ">Figure 12
<p>FE computed sensing coils peak voltage difference (with respect to torque = 0 Nm) versus applied pure torque.</p>
Full article ">Figure 13
<p>Sample shaft used to test the proposed system coupled to a 2 m long bar and a test rig.</p>
Full article ">Figure 14
<p>Experimental setup used to produce the static mechanical excitation.</p>
Full article ">Figure 15
<p>Sample shaft under test with mounted probehead (<b>a</b>) and 3D printed probehead support (<b>b</b>).</p>
Full article ">Figure 16
<p>Peak voltages on the sensing coils as a function of applied pure torque. Experimental results.</p>
Full article ">Figure 17
<p>Sensing coils peak voltage difference (with respect to null torque) versus applied pure torque. Experimental results.</p>
Full article ">Figure 18
<p>Sensing coils peak voltage difference (with respect to null load) versus applied torque in case of non-pure torque test. Experimental results.</p>
Full article ">Figure 19
<p>Sensing coils peak voltage versus applied torque: loading and unloading cycle. Experimental results.</p>
Full article ">Figure 20
<p>Sensing coils peak voltages versus yaw angles, in unloaded shaft conditions. Experimental results.</p>
Full article ">
12 pages, 2138 KiB  
Article
Unusual Anomalous Hall Effect in Two-Dimensional Ferromagnetic Cr7Te8
by Yifei Ma, Rui Yao, Jingrui Wu, Zhansheng Gao and Feng Luo
Molecules 2024, 29(21), 5068; https://doi.org/10.3390/molecules29215068 - 26 Oct 2024
Viewed by 713
Abstract
Two-dimensional (2D) materials with inherent magnetism have attracted considerable attention in the fields of spintronics and condensed matter physics. The anomalous Hall effect (AHE) offers a theoretical foundation for understanding the origins of 2D ferromagnetism (2D-FM) and offers a valuable opportunity for applications [...] Read more.
Two-dimensional (2D) materials with inherent magnetism have attracted considerable attention in the fields of spintronics and condensed matter physics. The anomalous Hall effect (AHE) offers a theoretical foundation for understanding the origins of 2D ferromagnetism (2D-FM) and offers a valuable opportunity for applications in topological electronics. Here, we present uniform and large-size 2D Cr7Te8 nanosheets with varying thicknesses grown using the chemical vapor deposition (CVD) method. The 2D Cr7Te8 nanosheets with robust perpendicular magnetic anisotropy, even a few layers deep, exhibit a Curie temperature (TC) ranging from 180 to 270 K according to the varying thickness of Cr7Te8. Moreover, we observed a temperature-induced reversal in the sign of the anomalous Hall resistance, correlating with changes in the intrinsic Berry curvature. Additionally, the topological Hall effect (THE) observed at low temperatures suggests the presence of non-trivial spin chirality. Our findings about topologically non-trivial magnetic spin states in 2D ferromagnets provide a promising opportunity for new designs in magnetic memory spintronics. Full article
(This article belongs to the Special Issue Novel Two-Dimensional Energy-Environmental Materials)
Show Figures

Figure 1

Figure 1
<p>Synthesis and characterization of 2D Cr<sub>7</sub>Te<sub>8</sub> nanosheets: (<b>a</b>) schematic diagram of the CVD system employed to grow Cr<sub>7</sub>Te<sub>8</sub> nanosheets, where Te and CrCl<sub>3</sub> sources were placed in separate zones with distinct heating processes; (<b>b</b>) typical optical microscopy (OM), illustrating the typical Cr<sub>7</sub>Te<sub>8</sub> nanosheets; (<b>c</b>) OM image illustrating a Cr<sub>7</sub>Te<sub>8</sub> nanosheet with an approximate domain size of 125 μm; (<b>d</b>) atomic force microscopy (AFM) image demonstrating the uniformity of the as-grown Cr<sub>7</sub>Te<sub>8</sub> nanosheets with a thickness of ~3.65 nm (<a href="#app1-molecules-29-05068" class="html-app">Figure S2</a>); (<b>e</b>) XRD pattern of CVD-grown Cr<sub>7</sub>Te<sub>8</sub> film.</p>
Full article ">Figure 2
<p>TEM characterization and atomic-scale structural analysis of the CVD-synthesized Cr<sub>7</sub>Te<sub>8</sub> nanosheet: (<b>a</b>–<b>c</b>) TEM images of a typical Cr<sub>7</sub>Te<sub>8</sub> nanosheet, oriented along the [001] zone axis and supported on a Cu grid, including the TEM image (<b>a</b>), SAED (<b>b</b>), and HRTEM (<b>c</b>), indicating that a hexagonal close-packed diffraction pattern was observed, and the lattice spacing of the (110) plane is about 2.0 Å. The yellow dashed rectangular box indicates the sampling location from which the measurements were obtained; (<b>d</b>) HRTEM image of the folded edge of a Cr<sub>7</sub>Te<sub>8</sub> nanosheet (see inset), revealing a distinct layered spacing of about 6.1 Å; (<b>e</b>,<b>f</b>) cross-sectional TEM images of Cr<sub>7</sub>Te<sub>8</sub> nanosheet on mica, aligned with [100] zone axis; (<b>e</b>) HAADF-STEM images display brighter contrast for Te (green ball) and darker contrast for Cr and Cr<sub>1</sub> (red ball and orange ball), consistent with the NiAs-type Cr<sub>7</sub>Te<sub>8</sub> crystal structure; (<b>f</b>) the corresponding ABF image.</p>
Full article ">Figure 3
<p>Electrical transport measurement of the CVD-grown Cr<sub>7</sub>Te<sub>8</sub> nanosheets: (<b>a</b>) Colored OM image of a three-dimensional Hall device. (<b>b</b>) Longitudinal resistance <span class="html-italic">R<sub>xx</sub></span>–<span class="html-italic">T</span> of a Cr<sub>7</sub>Te<sub>8</sub> (~20 nm) hall device. The upper-left inset depicts the temperature dependence of the first derivative of <span class="html-italic">R<sub>xx</sub></span>. (<b>c</b>) Temperature-dependent of Hall resistance (<span class="html-italic">R<sub>xy</sub></span>) loops. The presence of ferromagnetism is indicated by the rectangular hysteresis observed in <span class="html-italic">R<sub>xy</sub></span>. (<b>d</b>) The extracted <span class="html-italic">R<sub>xy</sub></span>–<span class="html-italic">μ</span><sub>0</sub><span class="html-italic">H</span> curve from 150 to 210 K, highlighting the Hall resistance reversal occurring between 170 K and 180 K. (<b>e</b>) The temperature dependence of <span class="html-italic">R</span><sub>H</sub>, <span class="html-italic">H</span><sub>C</sub>, and <span class="html-italic">R</span><sub>AHE(0)</sub> for the Cr<sub>7</sub>Te<sub>8</sub> hall device. (<b>f</b>) The Curie temperature <span class="html-italic">T</span><sub>C</sub> and transition temperature <span class="html-italic">T</span><sub>S</sub> of 2D Cr<sub>7</sub>Te<sub>8</sub> nanosheets devices with varying thicknesses were determined by analyzing the first derivative of the <span class="html-italic">R<sub>xx</sub></span>–<span class="html-italic">T</span> curve for <span class="html-italic">T</span><sub>C</sub> and examining the <span class="html-italic">R<sub>xy</sub></span>–<span class="html-italic">T</span> loops for <span class="html-italic">T</span><sub>S</sub>, identifying sharp slope changes and transition points, respectively. The blue and orange areas show the trend of <span class="html-italic">T</span><sub>C</sub> and <span class="html-italic">T</span><sub>S</sub> gradually increasing with the increase of thickness.</p>
Full article ">Figure 4
<p>Analysis of the temperature-dependent sign reversal of the <span class="html-italic">R<sub>xy</sub></span>. (<b>a</b>) Plot of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi mathvariant="normal">A</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">E</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </msub> </mrow> </semantics></math> versus the <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math> values at zero magnetic field for Cr<sub>7</sub>Te<sub>8</sub> hall devices of varying thicknesses and temperatures. The lines represent linear fits demonstrating the proportional relationship <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi mathvariant="normal">A</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">E</mi> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </msub> <mo>∝</mo> <mtext> </mtext> <msubsup> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> <mi>x</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msubsup> </mrow> </semantics></math>. (<b>b</b>) Plot of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> <mi>y</mi> </mrow> </msub> </mrow> </semantics></math> versus <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>σ</mi> </mrow> <mrow> <mi>x</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> for Cr<sub>7</sub>Te<sub>8</sub> nanosheets spanning the various AHE regimes from the impurity scattering mechanism (orange area) through the intrinsic (green area) and extrinsic regimes (blue area). The data points for all our Cr<sub>7</sub>Te<sub>8</sub> samples, regardless of thickness, fall within the intrinsic region.</p>
Full article ">Figure 5
<p>Investigation of potential topological hall effect in ultrathin Cr<sub>7</sub>Te<sub>8</sub> nanosheet: (<b>a</b>) A detailed analysis of the Hall resistivity for the Cr<sub>7</sub>Te<sub>8</sub> device (~4.3 nm) at <span class="html-italic">T</span> = 50 K with the subtraction of the OHE term. AHE term (indicated by the black lines) and THE term (highlighted by the green area) were shown. The AHE component is fitted by a <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>M</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mi mathvariant="normal">t</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">n</mi> <mi mathvariant="normal">h</mi> <mo>⁡</mo> <mo>(</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>H</mi> </mrow> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </mfrac> </mstyle> <mo>−</mo> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math> function, where <span class="html-italic">M</span><sub>0</sub>, <span class="html-italic">a</span><sub>0</sub>, and <span class="html-italic">H</span><sub>0</sub> are the fitting parameters. The inset is the OM image of the Cr<sub>7</sub>Te<sub>8</sub> (~4.3 nm) hall device. (<b>b</b>) Magnetic field dependence of the subtracted transverse resistance from 55 to 74 K, clearly showing the <span class="html-italic">T</span><sub>S</sub> between 62 and 66 K. (<b>c</b>) Temperature dependence of <span class="html-italic">R<sub>xx</sub></span> for the Cr<sub>7</sub>Te<sub>8</sub> hall device, showing a pronounced hump feature up to 74 K (highlighted by olive filled areas), which correlates closely with the presence of THE. (<b>d</b>,<b>e</b>) Typical field dependence curves of the <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">E</mi> </mrow> </msub> </mrow> </semantics></math> at various temperatures (40~74 K). (<b>f</b>) Temperature dependence of the <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">E</mi> </mrow> <mrow> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">x</mi> </mrow> </msubsup> </mrow> </semantics></math> (black line), along with the corresponding <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>H</mi> </mrow> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">H</mi> <mi mathvariant="normal">E</mi> </mrow> <mrow> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">x</mi> </mrow> </msubsup> </mrow> </semantics></math> (blue line) and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>H</mi> </mrow> <mrow> <mi mathvariant="normal">C</mi> </mrow> </msub> </mrow> </semantics></math> (red line). All the different colored arrows indicate the magnetic field sweep directions.</p>
Full article ">
14 pages, 1486 KiB  
Article
Analysis of Magnetization Dynamics in NiFe Thin Films with Growth-Induced Magnetic Anisotropies
by Leah Merryweather and Aidan T. Hindmarch
Magnetochemistry 2024, 10(10), 80; https://doi.org/10.3390/magnetochemistry10100080 - 21 Oct 2024
Viewed by 577
Abstract
We have used angled magnetron sputter deposition with and without sample rotation to control the magnetic anisotropy in 20 nm NiFe films. Ferromagnetic resonance spectroscopy, with data analysis using a Bayesian approach, is used to extract material parameters relating to the magnetic anisotropy. [...] Read more.
We have used angled magnetron sputter deposition with and without sample rotation to control the magnetic anisotropy in 20 nm NiFe films. Ferromagnetic resonance spectroscopy, with data analysis using a Bayesian approach, is used to extract material parameters relating to the magnetic anisotropy. When the sample is rotated during growth, only shape anisotropy is present, but when the sample is held fixed, a strong uniaxial anisotropy emerges with in-plane easy axis along the azimuthal direction of the incident atom flux. When the film is deposited in two steps, with an in-plane rotation of 90 degrees between steps, the two orthogonal induced in-plane easy-axes effectively cancel. The analysis approach enables precise and accurate determination of material parameters from ferromagnetic resonance measurements; this demonstrates the ability to precisely control both the direction and strength of uniaxial magnetic anisotropy, which is important in magnetic thin-film device applications. Full article
(This article belongs to the Special Issue Fabrication, Characterization and Application of Magnetic Thin Films)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Measured X-ray reflectivity for single layer NiFe films where the sample is held static during growth (upper trace, blue) and rotated about the sample normal at 10 rpm (lower trace, red). Best fitting models of the sample structure are shown as the solid lines through the data, which are generated from the structural scattering-length densities (sSLD) shown in (<b>b</b>). (<b>c</b>,<b>d</b>) show the same for the case where the sample is deposited in two steps, with the sample static during each step, but rotated by 90 degrees between steps. A slight variation in density is apparent across the ‘interface’ corresponding to this break in deposition.</p>
Full article ">Figure 2
<p>(<b>a</b>) Measured transmitted power signals (black points) and respective fits (red lines) showing the in-plane angular-dependence of the ferromagnetic resonance field at 10 GHz for ‘static’ sample with growth-induced uniaxial in-plane magnetic anisotropy. (<b>b</b>) Exemplar background-subtracted transmitted power signal and lineshape fit for signal at <math display="inline"><semantics> <mrow> <mi>θ</mi> <mo>=</mo> <mn>30</mn> </mrow> </semantics></math> degrees, corresponding roughly to magnetic field applied along the uniaxial hard-axis, and showing the absorption and dispersion components. Normalized residuals are shown in the lower panel. Best fit and residuals are shown in black, with red lines and markers showing the range of plausible solutions and associated residuals. (<b>c</b>) Histograms showing the posterior probability distributions of the parameters for the fit shown in (<b>b</b>). (<b>d</b>) Heatmaps of parameter pairs from the sampled posterior probability distributions showing correlation between pairs of fitting parameters.</p>
Full article ">Figure 3
<p>Angular dependence of the in-plane ferromagnetic resonance frequency for samples where (<b>a</b>) the sample is rotated during deposition, (<b>b</b>) the sample is held static during deposition, with the incident atom flux along an azimuthal angle corresponding to ∼120 degrees, and (<b>c</b>) the deposition is two-step, with the incident atom flux along ∼30 degrees in the first step and ∼120 degrees in the second step. Data and fits are shown with normalized residuals (below) on the right, and as polar plots on the left. The radial scale on the polar plots matches the vertical scale of the respective plots on the right. Solid black lines (open points in residual plots) correspond to the best fit (best fit residuals) and the broader overlapping red lines (filled points in the residual plots) show the range of feasible solutions (normalized residuals) sampled across posterior probability distributions.</p>
Full article ">Figure 4
<p>(<b>a</b>) Example ferromagnetic resonance measurement at 10 GHz for two-step sample, showing a single resonance peak, and resulting (<b>b</b>) parameter histograms and (<b>c</b>) correlation heatmaps. Best fit and residuals in (<b>a</b>) are shown in black, with red lines and markers showing the range of plausible solutions and associated residuals.</p>
Full article ">Figure 5
<p>(<b>a</b>) Measured power signals (black) and respective fits (red) for out-of-plane magnetization over a frequency range 5–20 GHz for the rotated sample. (<b>b</b>) Background-subtracted power signal and lineshape fit for signal at <math display="inline"><semantics> <mrow> <mi>F</mi> <mo>=</mo> <mn>10</mn> </mrow> </semantics></math> GHz, showing the absorption and dispersion components. Normalized residuals are shown in the lower panel. Best fit and residuals are shown in black, with red lines and markers showing the range of plausible solutions and associated residuals. (<b>c</b>) Histograms showing the posterior probability distributions of the parameters for the fit shown in (<b>b</b>), and (<b>d</b>) heatmaps of the sampled posterior probability distributions showing correlation between pairs of fitting parameters.</p>
Full article ">Figure 6
<p>Kittel fits—frequency-dependence of the resonance field—for sample held static during deposition to induce uniaxial in-plane magnetic anisotropy. In (<b>a</b>), the applied field is perpendicular to the plane of the film, whereas in (<b>b</b>,<b>c</b>), the magnetic field is applied at a pair of arbitrary orthogonal directions in the sample plane, i.e., not specifically along uniaxial easy/hard axes, for example. On the right of each frame are normalized residuals; the line and markers are as in <a href="#magnetochemistry-10-00080-f003" class="html-fig">Figure 3</a>.</p>
Full article ">
14 pages, 11980 KiB  
Article
Suitable Method for Improving Friction Performance of Magnetic Wheels with Metal Yokes
by Masaru Tanida, Kosuke Ono, Takehiro Shiba and Yogo Takada
Robotics 2024, 13(10), 151; https://doi.org/10.3390/robotics13100151 - 11 Oct 2024
Viewed by 569
Abstract
A magnetic-wheeled robot is a type of robot that inspects large steel structures instead of humans, and it can run on a three-dimensional path by using wheels with built-in permanent magnets. For the robots to work safely, their magnetic wheels require both magnetic [...] Read more.
A magnetic-wheeled robot is a type of robot that inspects large steel structures instead of humans, and it can run on a three-dimensional path by using wheels with built-in permanent magnets. For the robots to work safely, their magnetic wheels require both magnetic attractive forces and friction forces. Planetary-geared magnetic wheels, which we have developed, make direct contact with their yokes on the running surface to ensure their magnetic attractive force. However, this design decreases their frictional performance more than common magnetic wheels covered with soft materials. Therefore, the yokes require methods that can improve their frictional performance without decreasing their attractive force. To consider the best method for the use of magnetic wheels, this study has run experiments with five types of yokes, which have different processing. As a result, the yokes with corroded surfaces could have maintained the attractive force more than 90% of the time and increased their traction forces by about 36% in static conditions and about 30% in dynamic conditions compared to yokes with no machining. The main reasons for these experimental results are that the rust layer has stable irregularities on the surface and includes ferromagnetic materials. Full article
(This article belongs to the Section Intelligent Robots and Mechatronics)
Show Figures

Figure 1

Figure 1
<p>A planetary geared magnetic wheel (PGMW). (<b>a</b>) Front view. (<b>b</b>) Rear view.</p>
Full article ">Figure 2
<p>Inside of the PGMW.</p>
Full article ">Figure 3
<p>SCPREM-I. (<b>a</b>) Front left view. (<b>b</b>) Running through a flange path.</p>
Full article ">Figure 4
<p>The magnetic circuit formed in the PGMW.</p>
Full article ">Figure 5
<p>A newly designed PGMW.</p>
Full article ">Figure 6
<p>A towing robot. (<b>a</b>) Front left view. (<b>b</b>) Rear right view.</p>
Full article ">Figure 7
<p>Experimental apparatus.</p>
Full article ">Figure 8
<p>A bar chart showing normal force on the yokes’ two wheels with magnets <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>N</mi> </mrow> <mrow> <mn>2</mn> <mi>W</mi> </mrow> </msub> </mrow> </semantics></math>. Parenthetic percentage figures on the chart are rates with Yoke A as 100%.</p>
Full article ">Figure 9
<p>A bar chart showing the traction forces of the towing robot with magnets. Blue bars show static condition <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>. Orange bars show dynamic condition <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>. Parenthetic percentage figures on the chart are rates with Yoke A as 100%.</p>
Full article ">Figure 10
<p>A bar chart showing the yokes’ friction coefficients. The blue bars show the coefficients in a static condition <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>S</mi> </mrow> </msub> </mrow> </semantics></math>. The orange bars show the coefficients in dynamic condition <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>μ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>. Gray bars show the reduction rate of the coefficient <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>E</mi> </mrow> <mrow> <mi>μ</mi> </mrow> </msub> </mrow> </semantics></math>. The parenthetic percentage figures are the rates with Yoke A as 100%.</p>
Full article ">Figure 11
<p>Shifting of traction forces of Yoke E with magnets in dynamic conditions.</p>
Full article ">
19 pages, 5942 KiB  
Article
Research on Pipeline Stress Detection Method Based on Double Magnetic Coupling Technology
by Guoqing Wang, Qi Xia, Hong Yan, Shicheng Bei, Huakai Zhang, Hao Geng and Yuhan Zhao
Sensors 2024, 24(19), 6463; https://doi.org/10.3390/s24196463 - 7 Oct 2024
Viewed by 709
Abstract
Oil and gas pipelines are subject to soil corrosion and medium pressure factors, resulting in stress concentration and pipe rupture and explosion. Non-destructive testing technology can identify the stress concentration and defect corrosion area of the pipeline to ensure the safety of pipeline [...] Read more.
Oil and gas pipelines are subject to soil corrosion and medium pressure factors, resulting in stress concentration and pipe rupture and explosion. Non-destructive testing technology can identify the stress concentration and defect corrosion area of the pipeline to ensure the safety of pipeline transportation. In view of the problem that the traditional pipeline inspection cannot identify the stress signal at the defect, this paper proposes a detection method using strong and weak magnetic coupling technology. Based on the traditional J-A (Jiles–Atherton) model, the pinning coefficient is optimized and the stress demagnetization factor is added to establish the defect of the ferromagnetic material. The force-magnetic relationship optimization model is used to calculate the best detection magnetic field strength. The force-magnetic coupling simulation of Q235 steel material is carried out by ANSYS 2019 R1 software based on the improved J-A force-magnetic model. The results show that the effect of the stress on the pipe on the magnetic induction increases first and then decreases with the increase in the excitation magnetic field strength, and the magnetic signal has the maximum proportion of the stress signal during the excitation process; the magnetic induction at the pipe defect increases linearly with the increase in the stress trend. Through the strong and weak magnetic scanning detection of cracked pipeline materials, the correctness of the theoretical analysis and the validity of the engineering application of the strong and weak magnetic detection method are verified. Full article
Show Figures

Figure 1

Figure 1
<p>Magnetisation curves under different stresses.</p>
Full article ">Figure 2
<p>Schematic of the optimal detection of magnetisation.</p>
Full article ">Figure 3
<p>Change curve of the stress generation signal and applied magnetic field.</p>
Full article ">Figure 4
<p>Relationship curve between magnetic field intensity and magnetic induction intensity change rate.</p>
Full article ">Figure 5
<p>Defect steel mesh division model.</p>
Full article ">Figure 6
<p>Stress cloud diagram of steel defects.</p>
Full article ">Figure 7
<p>Axial composite signal.</p>
Full article ">Figure 8
<p>Radial composite signal.</p>
Full article ">Figure 9
<p>Proportion curve of the radial stress signal.</p>
Full article ">Figure 10
<p>Eigenvalues of composite signals under different stresses.</p>
Full article ">Figure 11
<p>Strong and weak magnetic scanning test platform.</p>
Full article ">Figure 12
<p>Strong and weak magnetic detector probe.</p>
Full article ">Figure 13
<p>Magnetic signal of the axial component detected by scanning.</p>
Full article ">Figure 14
<p>Magnetic signal of the radial component detected by scanning.</p>
Full article ">Figure 15
<p>Ratio of the stress signal and variation in the applied magnetic field.</p>
Full article ">Figure 16
<p>Relationship between the magnetic signal and stress.</p>
Full article ">
14 pages, 3952 KiB  
Article
Investigating Layered Topological Magnetic Materials as Efficient Electrocatalysts for the Hydrogen Evolution Reaction under High Current Densities
by Sanju Gupta, Hanna Świątek, Mirosław Sawczak, Tomasz Klimczuk and Robert Bogdanowicz
Catalysts 2024, 14(10), 676; https://doi.org/10.3390/catal14100676 - 1 Oct 2024
Viewed by 691
Abstract
Despite considerable progress, high-performing durable catalysts operating under large current densities (i.e., >1000 mA/cm2) are still lacking. To discover platinum group metal-free (PGM-free) electrocatalysts for sustainable energy, our research involves investigating layered topological magnetic materials (semiconducting ferromagnets) as highly efficient electrocatalysts [...] Read more.
Despite considerable progress, high-performing durable catalysts operating under large current densities (i.e., >1000 mA/cm2) are still lacking. To discover platinum group metal-free (PGM-free) electrocatalysts for sustainable energy, our research involves investigating layered topological magnetic materials (semiconducting ferromagnets) as highly efficient electrocatalysts for the hydrogen evolution reaction under high current densities and establishes the novel relations between structure and electrochemical property mechanisms. The materials of interest include transition metal trihalides, i.e., CrCl3, VCl3, and VI3, wherein a structural unit, the layered structure, is formed by Cr (or V) atoms sandwiched between two halides (Cl or I), forming a tri-layer. A few layers of quantum crystals were exfoliated (~50−60 nm), encapsulated with graphene, and electrocatalytic HER tests were conducted in acid (0.5M H2SO4) and alkaline (1M KOH) electrolytes. We find a reasonable HER activity evolved requiring overpotentials in a range of 30–50 mV under 10 mA cm−2 and 400−510 mV (0.5M H2SO4) and 280−500 mV (1M KOH) under −1000 mA cm−2. Likewise, the Tafel slopes range from 27 to 36 mV dec−1 (Volmer–Tafel) and 110 to 190 mV dec−1 (Volmer–Herovsky), implying that these mechanisms work at low and high current densities, respectively. Weak interlayer coupling, spontaneous surface oxidation, the presence of a semi-oxide subsurface (e.g., O–CrCl3), intrinsic Cl (or I) vacancy defects giving rise to in-gap states, electron redistribution (orbital hybridization) affecting the covalency, and sufficiently conductive support interaction lowering the charge transfer resistance endow the optimized adsorption/desorption strength of H* on active sites and favorable electrocatalytic properties. Such behavior is expedited for bi-/tri-layers while exemplifying the critical role of quantum nature electrocatalysts with defect sites for industrial-relevant conditions. Full article
(This article belongs to the Section Catalysis for Sustainable Energy)
Show Figures

Figure 1

Figure 1
<p><b>Morphology and structure.</b> (<b>a</b>–<b>d</b>) Optical photographs; (<b>e</b>–<b>h</b>) scanning electron micrographs; (<b>i</b>) θ–2θ XRD patterns; (<b>j</b>) Micro-Raman spectra excited with 514 nm of few-layered CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub> crystals; and (<b>k</b>) room-temperature electrical conductivity comparison with other reported HER catalysts. Also provided are conventional unit cells of MX<sub>3</sub> crystals with their respective colored atoms, Rietveld refinement, JCPDS nos., and an SEM image of substrate CNW/p–Si (001).</p>
Full article ">Figure 2
<p><b>Cyclic voltammograms and HER polarization curves.</b> (<b>a</b>,<b>b</b>) CV profiles, (<b>c</b>) capacitive current density difference versus scan rate, and (<b>d</b>,<b>e</b>) LSV polarization curves plotted in a large overpotential range with (dotted) and without (solid) considering <span class="html-italic">iR</span> ohmic drop of few-layer CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub> crystals, in acidic (0.5M H<sub>2</sub>SO<sub>4</sub>) and alkaline (1M KOH) electrolytes.</p>
Full article ">Figure 3
<p><b>Overpotential and Tafel slopes under low and high current densities.</b> Overpotential versus log current density, |j|, and Tafel slopes under low current densities of few-layer CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub>, in (<b>a</b>) acidic (0.5M H<sub>2</sub>SO<sub>4</sub>) and (<b>b</b>) alkaline (1M KOH) electrolytes. (<b>c</b>) Corresponding Tafel slope analysis under high current densities. The dotted curve shows the coverage-dependent current densities.</p>
Full article ">Figure 4
<p><b>Performance evaluation.</b> Comparison of the required overpotentials to reach the current density of 200 mA cm<sup>−2</sup> and 1000 mA cm<sup>−2</sup> between the recently reported catalysts to those studied here. The scale bar of the catalysts is based on the LSV measurements.</p>
Full article ">Figure 5
<p><b>Proposed HER mechanism in relation to the defects and electronic structure.</b> Illustrations of (<b>a</b>) the experimental scheme for high efficiency HER (<b>b</b>,<b>c</b>), the two-dimensional view of the crystal structure for MX<sub>3</sub> viewed along the (<b>b</b>) <span class="html-italic">a</span>-axis (planar) and (<b>c</b>) <span class="html-italic">c</span>-axis (side view), where the M (=Cr, V) atoms are the bigger spheres, and the X (Cl, I) atoms are the smaller spheres. The MX<sub>6</sub> octahedra form a layered honeycomb lattice via edge-sharing within each layer, and the layers are stacked in an ABC sequence along the <span class="html-italic">c</span>-axis. The bond lengths for CrCl<sub>3</sub> are shown in panel (<b>b</b>). Also provided are the presence of the Cl vacancy (dotted circle) and the vacant interstitial sites occupied by oxygen (solid red circle) in the honeycomb array responsible for the HER reaction. (<b>d</b>) Cartoons of a momentum space diagram and the DOS (density of states) for topological magnetic materials with the coexistence of the WSM and DNL behaviors of the surface/edge states.</p>
Full article ">
Back to TopTop