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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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
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
remove_circle_outline

Search Results (1,760)

Search Parameters:
Keywords = induced voltage

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 18318 KiB  
Article
Performance Analysis of a Synchronous Reluctance Generator with a Slitted-Rotor Core for Off-Grid Wind Power Generation
by Samuel Adjei-Frimpong and Mbika Muteba
Electricity 2025, 6(1), 2; https://doi.org/10.3390/electricity6010002 - 8 Jan 2025
Abstract
In this paper, the performance of a Dual-Stator Winding Synchronous Reluctance Generator (SynRG) suitability for off-grid wind power generation is analyzed. The rotor of the SynRG has a slitted-rotor core to improve selected vital performance parameters. The SynRG with a slitted-rotor core was [...] Read more.
In this paper, the performance of a Dual-Stator Winding Synchronous Reluctance Generator (SynRG) suitability for off-grid wind power generation is analyzed. The rotor of the SynRG has a slitted-rotor core to improve selected vital performance parameters. The SynRG with a slitted-rotor core was modeled using a two-dimensional (2D) Finite Element Method (FEM) to study the electromagnetic performance of key parameters of interest. To validate the FEA results, a prototype of the SynRG with a slitted rotor was tested in the laboratory for no-load operation and load operation for unity, lagging, and leading power factors. To evaluate the capability of the SynRG with a slitted-rotor core to operate in a wind turbine environment, the machine was modeled and simulated in Matlab/Simulink (R2023a) for dynamic responses. The FEA results reveal that the SynRG with a slitted-rotor core, compared with the conventional SynRG with the same ratings and specifications, reduces the torque ripple by 24.51%, 29.72%, and 13.13% when feeding 8 A to a load with unity, lagging, and leading power factors, respectively. The FEA results also show that the induced voltage on no-load of the SynRG with a slitted-rotor core, compared with the conventional SynRG of the same ratings and specifications, increases by 10.77% when the auxiliary winding is fed by a capacitive excitation current of 6 A. Furthermore, the same results show that with a fixed excitation capacitive current of 6 A, the effect of armature reaction of the SynRG with a slitted-rotor core is demagnetizing when operating with load currents having a lagging power factor, and magnetizing when operating with load currents having unity and leading power factors. The same patterns have been observed in the experimental results for different excitation capacitance values. The Matlab/Simulink results show that the SynRG with a slitted-rotor core has a quicker dynamic response than the conventional SynRG. However, a well-designed pitch-control mechanism for the wind turbine is necessary to account for changes in wind speeds. Full article
Show Figures

Figure 1

Figure 1
<p>Synchronous Reluctance Generator with direct capacitance injection. (<b>a</b>) Off-grid wind turbine topology, (<b>b</b>) Schematic diagram of the self-excited NSynRG winding scheme.</p>
Full article ">Figure 2
<p>Cross-sections of the rotor of SynRG and stator with two sets of windings. (<b>a</b>) Rotor with slit-cuts on <span class="html-italic">d</span>-axis, (<b>b</b>) Conventional rotor, (<b>c</b>) Stator showing the arrangement of both main and auxiliary windings.</p>
Full article ">Figure 3
<p>Photographs that show the constituents of the prototype SynRG with a slitted-rotor core. (<b>a</b>) Single rotor lamination, (<b>b</b>) Complete rotor core with shaft, (<b>c</b>) Rotor core inserted in the stator bore space ready to be fastened.</p>
Full article ">Figure 4
<p>Photographs of the laboratory setup. (<b>a</b>) Experimental setup rig photo, (<b>b</b>) Photograph showing the inductive load scheme in the experimental setup.</p>
Full article ">Figure 5
<p>Flux density distribution of the SynRG with a slitted-rotor core operating with a current load of 8 A and an excitation current of 6.8 A: 1000 rpm: (<b>a</b>) Unity power factor, (<b>b</b>) Lagging power factor, (<b>c</b>) Leading power factor.</p>
Full article ">Figure 6
<p>Comparison of saturation characteristics between the unoptimized NSynRG with a slitted-rotor core and optimized NSynRG with a slitted-rotor core.</p>
Full article ">Figure 7
<p>Induced no-load voltages with excitation current of 6 A. (<b>a</b>) Conventional SynRG, (<b>b</b>) SynRG with a slitted-rotor core.</p>
Full article ">Figure 8
<p>Airgap flux density on no-load with excitation current of 6.8 A. (<b>a</b>) The profile, (<b>b</b>) Harmonic spectrum.</p>
Full article ">Figure 9
<p>FEA electromagnetic torque profiles with a load current of 8 A of the NSynRG. (<b>a</b>) Unoptimized NSynRG, (<b>b</b>) Optimized NSynRG.</p>
Full article ">Figure 10
<p>Generated voltage characteristics on no-load. (<b>a</b>) Excitation capacitance of 50 µF, (<b>b</b>) Excitation capacitance of 90 µF, (<b>c</b>) Excitation capacitance of 140 µF, (<b>d</b>) Harmonic components of the no-load induced voltage.</p>
Full article ">Figure 11
<p>Regulation characteristics at different excitation capacitance values, such as (<b>a</b>) resistive load, (<b>b</b>) inductive load, and (<b>c</b>) capacitive load.</p>
Full article ">Figure 12
<p>No-load dynamic response of the SynRG with a slitted-rotor core at a constant wind speed of 4 m/s with 90 µF excitation capacitance. (<b>a</b>) Electromagnetic Torque, (<b>b</b>) Speed.</p>
Full article ">Figure 13
<p>No-load dynamic response of the conventional SynRG rotor at a constant wind speed of 4 m/s with 90 µF excitation capacitance. (<b>a</b>) Electromagnetic Torque, (<b>b</b>) Speed.</p>
Full article ">
12 pages, 3122 KiB  
Article
Effect of p-InGaN Cap Layer on Low-Resistance Contact to p-GaN: Carrier Transport Mechanism and Barrier Height Characteristics
by Mohit Kumar, Laurent Xu, Timothée Labau, Jérôme Biscarrat, Simona Torrengo, Matthew Charles, Christophe Lecouvey, Aurélien Olivier, Joelle Zgheib, René Escoffier and Julien Buckley
Crystals 2025, 15(1), 56; https://doi.org/10.3390/cryst15010056 - 8 Jan 2025
Viewed by 208
Abstract
This study investigated the low contact resistivity and Schottky barrier characteristics in p-GaN by modifying the thickness and doping levels of a p-InGaN cap layer. A comparative analysis with highly doped p-InGaN revealed the key mechanisms contributing to low-resistance contacts. Atomic force microscopy [...] Read more.
This study investigated the low contact resistivity and Schottky barrier characteristics in p-GaN by modifying the thickness and doping levels of a p-InGaN cap layer. A comparative analysis with highly doped p-InGaN revealed the key mechanisms contributing to low-resistance contacts. Atomic force microscopy inspections showed that the surface roughness depends on the doping levels and cap layer thickness, with higher doping improving the surface quality. Notably, increasing the doping concentration in the p++-InGaN cap layer significantly reduced the specific contact resistivity to 6.4 ± 0.8 × 10−6 Ω·cm2, primarily through enhanced tunneling. Current–voltage (I–V) characteristics indicated that the cap layer’s surface properties and strain-induced polarization effects influenced the Schottky barrier height and reverse current. The reduction in barrier height by approximately 0.42 eV in the p++-InGaN layer enhanced hole tunneling, further lowering the contact resistivity. Additionally, polarization-induced free charges at the metal–semiconductor interface reduced band bending, thereby enhancing carrier transport. A transition in current conduction mechanisms was also observed, shifting from recombination tunneling to space-charge-limited conduction across different voltage ranges. This research underscores the importance of doping, cap layer thickness, and polarization effects in achieving ultra-low contact resistivity, offering valuable insights for improving the performance of p-GaN-based power devices. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the epitaxial structure of the device, consisting of the following layers: (<b>a</b>) p<sup>+</sup>-GaN top layer, (<b>b</b>) p<sup>++</sup>-GaN heavily doped cap layer, (<b>c</b>) p<sup>++</sup>-In<sub>0.15</sub>Ga<sub>0.85</sub>N heavily doped cap layer, (<b>d</b>) etched circular pad stack, and (<b>e</b>) circular transmission line model (CTLM) configuration with 1, 2, 4, 8, 19, and 49 μm spacing.</p>
Full article ">Figure 2
<p>AFM image of samples L.1, L.2, M., N.1, N.2, N.3, and O.</p>
Full article ">Figure 3
<p>Room temperature current−voltage (I−V) characteristics of (<b>a</b>) samples L.1, L.2, and M., (<b>b</b>) samples N.1, N.2, N.3, and O., and (<b>c</b>) the analysis of the current conduction mechanism under forward bias for there difference voltage range named region 1, 2 and 3 for the p<sup>++</sup>-InGaN/p<sup>++</sup>-GaN heterojunction in sample O.</p>
Full article ">Figure 4
<p>Plot of the specific contact resistance for various samples. The contact transfer length method (CTLM) was employed, where a logarithmic fit to the experimental data was used to determine the specific contact resistance values.</p>
Full article ">Figure 5
<p>Current−voltage (I−V) curves and corresponding specific contact resistance measurements at the linear regime close to 15 V before annealing (<b>a</b>,<b>b</b>) and after annealing at 760 °C in an N<sub>2</sub> atmosphere (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Plot of the Schottky barrier height (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>Φ</mi> </mrow> <mrow> <mi>B</mi> </mrow> </msub> </mrow> </semantics></math>) for various samples, highlighting variations in barrier properties. (<b>b</b>) Energy band diagrams depicting (<b>i</b>) the metal/p<sup>+</sup>-GaN and (<b>ii</b>) the metal/p<sup>++</sup>-InGaN/p<sup>+</sup>-GaN superlattice structure.</p>
Full article ">
16 pages, 1709 KiB  
Article
An Optimized H5 Hysteresis Current Control with Clamped Diodes in Transformer-Less Grid-PV Inverter
by Sushil Phuyal, Shashwot Shrestha, Swodesh Sharma, Rachana Subedi, Anil Kumar Panjiyar and Mukesh Gautam
Electricity 2025, 6(1), 1; https://doi.org/10.3390/electricity6010001 - 7 Jan 2025
Viewed by 198
Abstract
With the rise of renewable energy penetration in the grid, photovoltaic (PV) panels are connected to the grid via inverters to supply solar energy. Transformer-less grid-tied PV inverters are gaining popularity because of their improved efficiency, reduced size, and lower costs. However, they [...] Read more.
With the rise of renewable energy penetration in the grid, photovoltaic (PV) panels are connected to the grid via inverters to supply solar energy. Transformer-less grid-tied PV inverters are gaining popularity because of their improved efficiency, reduced size, and lower costs. However, they can induce a path for leakage currents between the PV and the grid due to the absence of galvanic isolation. This leads to serious electromagnetic interference, loss in efficiency, and safety concerns. The leakage current is primarily influenced by the nature of the common mode voltage (CMV), which is determined by the switching techniques of the inverter. In this paper, a novel inverter topology of Hysteresis Controlled H5 with Two Clamping Diodes (HCH5-D2) is derived. The HCH5-D2 topology helps decouple the AC part (Grid) and DC part (PV) during the freewheeling period to make the CMV constant, thereby reducing the leakage current. Additionally, the extra diodes help reduce voltage spikes generated during the freewheeling period and maintain the CMV at a constant value. Finally, a 2.2 kW grid-connected single-phase HCH5-D2 PV inverter system’s MATLAB simulation is presented, showing better results compared to a traditional H4 inverter. Full article
Show Figures

Figure 1

Figure 1
<p>Main circuit of HCH5-D2 inverter topology.</p>
Full article ">Figure 2
<p>Resonant circuits of the H5 inverter. (<b>a</b>) Resonant circuit in terms of pole voltages. (<b>b</b>) Resonant circuit in terms of DM and CM voltages.</p>
Full article ">Figure 3
<p>Resonant circuit of H5 inverter with current sources.</p>
Full article ">Figure 4
<p>Resonant circuits of H5 inverter in various voltage and current source configurations. (<b>a</b>) A single voltage and current source. (<b>b</b>) In the form of two voltage sources.</p>
Full article ">Figure 5
<p>Simplified resonant circuit of H5 inverter in the form of single common-mode voltage source.</p>
Full article ">Figure 6
<p>Different modes of operation of HCH5D2. Mode (<b>1</b>). Positive power transfer. Mode (<b>2</b>). Positive freewheeling. Mode (<b>3</b>). Negative power transfer. Mode (<b>4</b>). Negative freewheeling.</p>
Full article ">Figure 7
<p>A control structure diagram for HCH5D2 inverter system, red arrow representing the five gating signals passed to the MOSFETs of the proposed inverter.</p>
Full article ">Figure 8
<p>A visual representation of hysteresis band current control.</p>
Full article ">Figure 9
<p>Simulation results of the HCH5-D2 topology showing the DC-link voltage and the corresponding switching pattern. The left plot (<b>a</b>) illustrates the stable DC-link voltage, while the right diagram (<b>b</b>) shows the switching pattern for different switches (S1–S5) over time.</p>
Full article ">Figure 10
<p>Simulation results of HCH5-D2 topology. Together, the graphs demonstrate the voltage behavior for both common and differential modes in the simulation of the clamped H5 inverter topology, reflecting stability in common mode and regular AC-like switching in differential mode.</p>
Full article ">Figure 11
<p>Comparison of simulation results: current. The HCH5-D2 topology shows almost no leakage current, indicating a more efficient and safer design compared to the fluctuating spikes of leakage current in the H4 topology.</p>
Full article ">Figure 12
<p>Comparative FFT analysis.</p>
Full article ">Figure 13
<p>Simulation results of the HCH5-D2 topology showing the grid voltage and injected grid current. The voltage and current waveforms are sinusoidal, indicating stable operation of the inverter.</p>
Full article ">Figure 14
<p>Experimental setup of HCH5D2 inverter topology.</p>
Full article ">
20 pages, 3388 KiB  
Article
Transcriptomic and Functional Landscape of Adult Human Spinal Cord NSPCs Compared to iPSC-Derived Neural Progenitor Cells
by Sasi Kumar Jagadeesan, Ahmad Galuta, Ryan Vimukthi Sandarage and Eve Chung Tsai
Cells 2025, 14(2), 64; https://doi.org/10.3390/cells14020064 - 7 Jan 2025
Viewed by 152
Abstract
The adult human spinal cord harbors diverse populations of neural stem/progenitor cells (NSPCs) essential for neuroregeneration and central nervous system repair. While induced pluripotent stem cell (iPSC)-derived NSPCs offer significant therapeutic potential, understanding their molecular and functional alignment with bona fide spinal cord [...] Read more.
The adult human spinal cord harbors diverse populations of neural stem/progenitor cells (NSPCs) essential for neuroregeneration and central nervous system repair. While induced pluripotent stem cell (iPSC)-derived NSPCs offer significant therapeutic potential, understanding their molecular and functional alignment with bona fide spinal cord NSPCs is crucial for developing autologous cell therapies that enhance spinal cord regeneration and minimize immune rejection. In this study, we present the first direct transcriptomic and functional comparison of syngeneic adult human NSPC populations, including bona fide spinal cord NSPCs and iPSC-derived NSPCs regionalized to the spinal cord (iPSC-SC) and forebrain (iPSC-Br). RNA sequencing analysis revealed distinct transcriptomic profiles and functional disparities among NSPC types. iPSC-Br NSPCs exhibited a close resemblance to bona fide spinal cord NSPCs, characterized by enriched expression of neurogenesis, axon guidance, synaptic signaling, and voltage-gated calcium channel activity pathways. Conversely, iPSC-SC NSPCs displayed significant heterogeneity, suboptimal regional specification, and elevated expression of neural crest and immune response-associated genes. Functional assays corroborated the transcriptomic findings, demonstrating superior neurogenic potential in iPSC-Br NSPCs. Additionally, we assessed donor-specific influences on NSPC behavior by analyzing gene expression and differentiation outcomes across syngeneic populations from multiple individuals. Donor-specific factors significantly modulated transcriptomic profiles, with notable variability in the alignment of iPSC-derived NSPCs to bona fide spinal cord NSPCs. Enrichment of pathways related to neurogenesis, axon guidance, and synaptic signaling varied across donors, highlighting the impact of genetic and epigenetic individuality on NSPC behavior. Full article
Show Figures

Figure 1

Figure 1
<p>Differentiation, proliferation, and characterization of NSPCs derived from iPSCs and bona fide source. iPSC-Br NSPCs more closely resemble bona fide NSPCs in morphology and marker expression compared to iPSC-SC NSPCs. (<b>a</b>) Schematic illustration of the experimental workflow. Skin tissue samples from human donors were cultured into primary fibroblasts and reprogrammed into iPSCs. (<b>b</b>) Morphological and immunocytochemical analysis of NSPCs. iPSC-SC NSPCs form flat and neurosphere-like colonies, expressing Sox2, Nestin, β-III tubulin, and Map2. iPSC-Br NSPCs exhibit multipolar spindle-like cells, expressing Sox2, Pax6, β-III tubulin, and GFAP. Bona fide spinal cord NSPCs form adherent layers and neurosphere-like colonies, expressing Nestin, Sox2, β-III tubulin, and GFAP. (<b>c</b>) Pluripotency markers (Sox2, Nanog, Oct3/4, Oct4a) in iPSC-SC colonies confirm retention of stemness. DAPI shows nuclei in all panels.</p>
Full article ">Figure 2
<p>Comparative analysis of differentiation and proliferation profiles between bona fide and iPSC-derived NSPC. iPSC-Br NSPCs showed greater neuronal and astrocyte differentiation and higher proliferation rates than both iPSC-SC and bona fide NSPCs, indicating their superior neurogenic potential. (<b>A</b>) The percentage of β-III tubulin+ neurons was significantly higher in bona fide, and iPSC-Br NSPCs compared to iPSC-SC NSPCs at both time points. (<b>B</b>) By week two, GFAP+ astrocyte differentiation was markedly higher in iPSC-Br NSPCs compared to bona fide and iPSC-SC NSPCs, with minimal astrocytic differentiation observed in the latter. (<b>C</b>) Proliferation rates, measured as BrdU+ cell percentages, were significantly higher in iPSC-derived NSPCs compared to bona fide NSPCs under differentiation conditions. (<b>D</b>) Lineage differentiation profiles revealed that iPSC-Br NSPCs exhibited contributions from neuronal, astrocytic, and minimal oligodendrocytic lineages, while iPSC-SC NSPCs were predominantly neuronal. Bona fide NSPCs were also predominantly neuronal with negligible astrocytic and almost no oligodendrocytic differentiation. (<b>E</b>) Self-renewal capacity, as indicated by BrdU+/Sox2+ cells, was higher in iPSC-derived NSPCs compared to bona fide NSPCs at both time points. Data are presented as mean ± s.e.m.; ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001; n = 3 for bona fide, iPSC-SC, and iPSC-Br NSPCs. Each assay was conducted with three biological replicates (H17, H18, H25), and each functional assay, such as the percentage of β-III tubulin + cells, was performed with three technical replicates. Each technical replicate involved capturing 10 standardized images from a 96-well plate.</p>
Full article ">Figure 3
<p>The transcriptomes of iPSC-Br and bona fide NSPCs are more similar than iPSC-SC and bona fide NSPCs. The gene expression analysis reveals that iPSC−Br NSPCs retain transcriptomic features similar to bona fide NSPCs, unlike the more distinct iPSC−SC NSPCs (<b>A</b>) Venn diagram of DE genes in both iPSC−SC and iPSC−Br NSPCs relative to bona fide NSPCs, including directionality (upregulation or downregulation). A total of 4600 genes were DE in both iPSC−SC and iPSC−Br NSPCs with a fold change ≥ 2 and q-value &lt; 0.05. (<b>B</b>) The top 30 genes DE in the same direction (upregulation or downregulation) in both iPSC−SC and iPSC−Br NSPCs relative to bona fide NSPCs. (<b>C</b>) PCA of the 500 most variable genes across bona fide, iPSC-SC, and iPSC-Br NSPCs. Each dot represents a sample (H17, H18, H25), with donor-specific identifiers labeled. PC1 (69% variance) explains the largest transcriptomic differences, while PC2 (7% variance) and PC3 (6% variance) capture additional variation. iPSC-Br NSPCs cluster closer to bona fide NSPCs than iPSC-SC NSPCs, reflecting greater transcriptomic similarity. Donor-specific differences are evident along PC2, while PC3 highlights intrinsic differences between iPSC-SC and iPSC-Br NSPCs. All differences are statistically significant, * <span class="html-italic">p</span> &lt; 0.05 for all comparisons. (<b>D</b>) Hierarchical clustering analysis for all transcripts (24,006 genes) found in common among bona fide, iPSC−SC, and iPSC−Br NSPCs. The clustering illustrates that bona fide NSPCs group closely together, with iPSC-Br NSPCs forming a cluster more similar to bona fide NSPCs than iPSC-SC NSPCs. The intensity of the green shading reflects expression levels, with darker shades indicating higher expression.</p>
Full article ">Figure 4
<p>Comparative gene expression analysis of iPSC-derived NSPCs and bona fide NSPCs. iPSC-SC NSPCs may reflect a distinct reprogramming pathway compared to iPSC-Br NSPCs. (<b>A</b>) Volcano plots show differentially expressed (DE) genes between iPSC-SC NSPCs (left panel) and iPSC-Br NSPCs (right panel) relative to bona fide NSPCs, with red dots indicating significant DE genes (fold change ≥ 2, <span class="html-italic">p</span>-value &lt; 0.05). (<b>B</b>) Heatmaps with hierarchical clustering depict gene expression patterns, comparing iPSC-SC NSPCs (left panel) and iPSC-Br NSPCs (right panel) to bona fide NSPCs. (<b>C</b>) Scatter plots highlight genes uniquely differentially expressed in iPSC-SC NSPCs (left panel) and iPSC-Br NSPCs (right panel) compared to bona fide NSPCs. (<b>D</b>) Lists of the top 50 upregulated and downregulated genes in iPSC-SC NSPCs and iPSC-Br NSPCs, each relative to bona fide NSPCs.</p>
Full article ">Figure 5
<p>Comparative gene set enrichment in iPSC-derived NSPCs relative to bona fide NSPCs. Gene set enrichment analysis was conducted using the GSEA Preranked algorithm against a subset of MSigDB gene sets. A normalized enrichment score (NES) was used to compare gene sets across NSPCs in the categories of (<b>A</b>) hallmark gene sets, (<b>B</b>) GO molecular functions, (<b>C</b>) canonical pathways, (<b>D</b>) GO biological processes, and (<b>E</b>) cell-type signatures. Gene sets with significant positive or negative NES are indicated by *, representing a q-value &lt; 0.05. iPSC-SC (n = 3) and iPSC-Brain (n = 3) are represented by pink and gray bars, respectively. Protein–protein interaction network maps of the top 30 (<b>F</b>) upregulated and (<b>G</b>) downregulated DE genes in both iPSC-SC and iPSC-Br NSPCs relative to bona fide NSPCs were generated using STRING v12.0 software (PPI enrichment <span class="html-italic">p</span>-value &lt; 1 × 10⁻¹⁶). Each node represents proteins from a single, protein-coding gene locus, with colored nodes indicating query proteins and first-shell interactions, and white nodes representing second-shell interactions. Edge confidence is proportional to line thickness, indicating the strength of protein–protein associations.</p>
Full article ">
17 pages, 13090 KiB  
Article
Dynamic Imaging of Projected Electric Potentials of Operando Semiconductor Devices by Time-Resolved Electron Holography
by Tolga Wagner, Hüseyin Çelik, Simon Gaebel, Dirk Berger, Peng-Han Lu, Ines Häusler, Nina Owschimikow, Michael Lehmann, Rafal E. Dunin-Borkowski, Christoph T. Koch and Fariba Hatami
Electronics 2025, 14(1), 199; https://doi.org/10.3390/electronics14010199 - 5 Jan 2025
Viewed by 522
Abstract
Interference gating (iGate) has emerged as a groundbreaking technique for ultrafast time-resolved electron holography in transmission electron microscopy, delivering nanometer spatial and nanosecond temporal resolution with minimal technological overhead. This study employs iGate to dynamically observe the local projected electric potential within the [...] Read more.
Interference gating (iGate) has emerged as a groundbreaking technique for ultrafast time-resolved electron holography in transmission electron microscopy, delivering nanometer spatial and nanosecond temporal resolution with minimal technological overhead. This study employs iGate to dynamically observe the local projected electric potential within the space-charge region of a contacted transmission electron microscopy (TEM) lamella manufactured from a silicon diode during switching between unbiased and reverse-biased conditions, achieving a temporal resolution of 25 ns at a repetition rate of 3 MHz. By synchronizing the holographic acquisition with the applied voltage, this approach enables the direct visualization of time-dependent potential distributions with high precision. Complementary static and dynamic experiments reveal a remarkable correspondence between modeled and measured projected potentials, validating the method’s robustness. The observed dynamic phase progressions resolve and allow one to differentiate between localized switching dynamics and preparation-induced effects, such as charge recombination near the sample edges. These results establish iGate as a transformative tool for operando investigations of semiconductor devices, paving the way for advancing the nanoscale imaging of high-speed electronic processes. Full article
(This article belongs to the Section Optoelectronics)
Show Figures

Figure 1

Figure 1
<p>Schematic of the time-resolved electron holography setup with interference gating in a transmission electron microscope (TEM). (<b>a</b>) The TEM configuration uses an RF biasing holder to apply a periodic voltage to the sample, creating an electron hologram by overlapping object (Obj) and reference (Ref) waves with a biprism. (<b>b</b>) Holographic reconstruction process: Fourier transformation (FT), isolating sideband (SB) from centerband (CB), extracting amplitude and phase information. (<b>c</b>) Interference Gating: dynamic fringe contrast <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>h</mi> <mi>o</mi> <mi>l</mi> </mrow> </msub> </semantics></math> with gate length <math display="inline"><semantics> <mi>τ</mi> </semantics></math> (top panel), FT within and outside the gate (second panel), noise-based gating signal applied to dynamic phase shifter (third panel), control signal applied to RF biasing holder (bottom panel), synchronized to each other with an adjustable temporal delay <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> </mrow> </semantics></math> for setting the gate position <math display="inline"><semantics> <msub> <mi>t</mi> <mi>g</mi> </msub> </semantics></math>.</p>
Full article ">Figure 2
<p>(<b>a</b>) Sample preparation and multiscale imaging of the UG1A diode, showcasing the individual steps, ranging from the mechanical preparation to the contacted TEM lamella. The top panel shows a macroscopic view (light microscopy, LM) of the mechanically ground UG1A diode with the device visible, centered in between p- and n-contacts. The middle panel displays a voltage contrast image acquired by Scanning Electron Microscopy (SEM) utilizing a micro-manipulator (colored red) as an electrical contact, indicating a potential difference across the p–n junction interface (plotted in red, from <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>2</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> on the p-side to 0 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> on the n-side). The bottom panels show SEM images of the FIB-prepared lamella in top (electron image) and side views (ion image), highlighting the p- and n-doped areas within and the vacuum region surrounding the sample. (<b>b</b>) Potential model using SIMP. The upper diagram depicts the initial 2D potential distribution across the p–n junction (red line) and within the supporting chip with an applied reverse-bias <math display="inline"><semantics> <msub> <mi>V</mi> <mrow> <mi>t</mi> <mi>o</mi> <mi>t</mi> </mrow> </msub> </semantics></math>, while the black dashed lines represent the electric potential extending into the vacuum calculated by SIMP. The lower diagram shows a schematic cross-section of the SIMP-based extension of the initial 2D potential with an effective thickness <math display="inline"><semantics> <msub> <mi>t</mi> <mrow> <mi>e</mi> <mi>f</mi> <mi>f</mi> </mrow> </msub> </semantics></math> along the <span class="html-italic">z</span>-axis to the full 3D potential distribution, needed for calculating the projected potentials and simulated phases.</p>
Full article ">Figure 3
<p>(<b>a</b>) Normalized simulated phase (calculated by SIMP) of the UG1A diode under reverse-bias condition (<math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mrow> <mi>e</mi> <mi>x</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> <mo> </mo> <mi mathvariant="normal">V</mi> </mrow> </semantics></math>). The white boxes indicate the areas used as reference and object wave regions in the static electron holographic experiments and the dashed white box highlight the area of the contacted TEM lamella. (<b>b</b>) Comparison of the modeled phase (difference between object and reference wave regions, top row) with experimental phase reconstructions (bottom row) at different applied biases: <math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mrow> <mi>e</mi> <mi>x</mi> <mi>t</mi> </mrow> </msub> <mo>=</mo> <mo>−</mo> <mn>2</mn> <mo> </mo> <mi mathvariant="normal">V</mi> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>, and 0 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>. The dashed white boxes highlight the area of the contacted TEM lamella. (<b>c</b>) Phase profiles extracted from SIMP (dotted lines) and experimental data (solid lines) across the diode for varying biasing conditions.</p>
Full article ">Figure 4
<p>(<b>a</b>) Static phase reconstruction of the UG1A diode biased with <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>, highlighting the space-charge region (SCR) between the p- and n-doped sides. The dashed white box outlines the area of the contacted TEM lamella; the orange polygon outlines the FoV for the time-resolved measurements. (<b>b</b>) Reconstructed dynamic phases, acquired with a time resolution of <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo>=</mo> <mn>25</mn> <mo> </mo> <mi>ns</mi> </mrow> </semantics></math> at a repetition rate of <math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>3</mn> <mo> </mo> <mi mathvariant="normal">MHz</mi> </mrow> </semantics></math>, showing the phase distribution for different switching states (<math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>150</mn> <mo> </mo> <mi>ns</mi> </mrow> </semantics></math> and 0 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> at <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>=</mo> <mn>61</mn> <mo> </mo> <mi>ns</mi> </mrow> </semantics></math>) of the diode. The regions marked by orange lines indicate the areas used for the phase profiles in <span class="html-italic">x</span>- and <span class="html-italic">y</span>-directions. (<b>c</b>) Phase profiles at different biases (<math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> and 0 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>) with orange dashed lines indicating the spatial range for time-resolved measurements (small FoV), (<b>d</b>) plot of the phase slopes along the <span class="html-italic">x</span>-axis, and (<b>e</b>) phase profiles along the <span class="html-italic">y</span>-axis for <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> and 0 <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Dynamic phase frame of the UG1A diode at <math display="inline"><semantics> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </semantics></math> <math display="inline"><semantics> <mi mathvariant="normal">V</mi> </semantics></math> bias (150 ns), depicting the SCR and its extension into the n-doped region, with the subregions (I–IV) marked for the analysis of the temporal phase progression (in (<b>b</b>)). The lower panel schematically illustrates the position of the equi-phase lines. (<b>b</b>) Normalized phase values over time, averaged in each subregion (I–IV) during diode switching, showing localized phase modulations corresponding to bias changes. The gray-shaded areas indicate the location-dependent transitions captured by iGate. For improved visibility, the period of <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>333.3</mn> <mo> </mo> <mi>ns</mi> </mrow> </semantics></math> is repeated (<math display="inline"><semantics> <mrow> <mi>f</mi> <mo>=</mo> <mn>3</mn> <mo> </mo> <mi>MHz</mi> </mrow> </semantics></math>). (<b>c</b>) Sequential phase frames of the switching behavior into and out of reverse-bias condition, demonstrating the temporal evolution of the equi-phase lines (red dashed) within the SCR.</p>
Full article ">
13 pages, 2429 KiB  
Article
Non-Steroidal Anti-Inflammatory Drugs Are Inhibitors of the Intestinal Proton-Coupled Amino Acid Transporter (PAT1): Ibuprofen and Diclofenac Are Non-Translocated Inhibitors
by Carsten Uhd Nielsen, Sebastian Jakobsen and Maria L. Pedersen
Pharmaceutics 2025, 17(1), 49; https://doi.org/10.3390/pharmaceutics17010049 - 2 Jan 2025
Viewed by 318
Abstract
Background/Objectives: The proton-coupled amino acid transporter (PAT1) is an intestinal absorptive solute carrier responsible for the oral bioavailability of some GABA-mimetic drug substances such as vigabatrin and gaboxadol. In the present work, we investigate if non-steroidal anti-inflammatory drug substances (NSAIDs) interact with [...] Read more.
Background/Objectives: The proton-coupled amino acid transporter (PAT1) is an intestinal absorptive solute carrier responsible for the oral bioavailability of some GABA-mimetic drug substances such as vigabatrin and gaboxadol. In the present work, we investigate if non-steroidal anti-inflammatory drug substances (NSAIDs) interact with substrate transport via human (h)PAT1. Methods: The transport of substrates via hPAT1 was investigated in Caco-2 cells using radiolabeled substrate uptake and in X. laevis oocytes injected with hPAT1 cRNA, measuring induced currents using the two-electrode voltage clamp technique. The molecular interaction between NSAIDs and hPAT1 was investigated using an AlphaFold2 model and molecular docking. Results: NSAIDs such as ibuprofen, diclofenac, and flurbiprofen inhibited proline uptake via hPAT1, with IC50 values of 954 (logIC50 2.98 ± 0.1) µM, 272 (logIC50 2.43 ± 0.1) µM, and 280 (logIC50 2.45 ± 0.1) µM, respectively. Ibuprofen acted as a non-competitive inhibitor of hPAT1-mediated proline transport. In hPAT1-expressing oocytes, ibuprofen and diclofenac did not induce inward currents, and inhibited inward currents caused by proline. Molecular modeling pointed to a binding mode involving an allosteric site. Conclusions: NSAIDs interact with hPAT1 as non-translocated non-competitive inhibitors, and molecular modeling points to a binding mode involving an allosteric site distinct from the substrate binding site. The present findings could be used as a starting point for developing specific hPAT1 inhibitors. Full article
(This article belongs to the Section Drug Targeting and Design)
Show Figures

Figure 1

Figure 1
<p>Chemical structure and names of the compounds investigated for hPAT1 interaction. ASA: Acetylsalicylic acid.</p>
Full article ">Figure 2
<p>Proline uptake in Caco-2 cells measured in the absence or presence of NSAIDs. (<b>A</b>). The uptake rate of 13.3 nM [<sup>3</sup>H]-proline (1.0 µCi ml<sup>−1</sup>) in 10 mM MES HBSS buffer, pH 6.0 (control), and the presence of various NSAIDs and paracetamol (n = 3–4). (<b>B</b>). LDH release (n = 3). Each value represents the mean ± SEM. (<b>C</b>). The uptake rate of 13.3 nM [<sup>3</sup>H]-proline (1.0 µCi mL<sup>−1</sup>) in 10 mM MES HBSS buffer, pH 6.0 (control) and the presence of indomethacin, diclofenac, and ibuprofen (n = 3). (<b>D</b>). The uptake rate of 13.3 nM [<sup>3</sup>H]-proline (1.0 µCi mL<sup>−1</sup>) in 10 mM HEPES HBSS buffer, pH 7.4 (control), and the presence of indomethacin, diclofenac, and ibuprofen (n = 3).</p>
Full article ">Figure 3
<p>Concentrations-dependent inhibition of 13.3 nM [<sup>3</sup>H]-proline (1.0 µCi mL<sup>−1</sup>) uptake rate (%) at pH 6.0 in Caco-2 cells of diclofenac, flurbiprofen, or ibuprofen. The IC<sub>50</sub> value was determined from Equation (2). Each value represents the mean ± SEM (n = 3).</p>
Full article ">Figure 4
<p>Concentration-dependent proline uptake in Caco-2 cells in the absence or presence of ibuprofen at pH 6.0. The uptake rate of 13.3 nM [<sup>3</sup>H]-proline (1.0 µCi mL<sup>−1</sup>) with 0–10 mM L-proline (o), in the presence of 0.5 mM ibuprofen (●), 1.0 mM ibuprofen (■), or 2.0 mM ibuprofen (▲). Each value represents the mean ± SEM (n = 3). The solid lines show the fit of the resulting data to Equation (1).</p>
Full article ">Figure 5
<p>Two-electrode voltage clamp (TEVC) measurements in PAT1-expressing <span class="html-italic">Xenopus Laevis</span> oocytes. (<b>A</b>). The traces are representative of experiments performed on 4–6 different oocytes. The upper trace is a trace in a water-injected oocyte, whereas the lower trace is in a oocyte injected with <span class="html-italic">SLC36A1 cRNA</span>. The induced current in PAT1-expressing oocytes at a holding potential of −60 mV and continuously perfused with Ringer’s solution pH 6.0. 1: 20.0 mM L-Pro; 2: 20 mM L-Pro with 1.0 mM ibuprofen; 3: 1.0 mM ibuprofen; 4: 20.0 mM L-Pro with 1.0 mM diclofenac; 5: 1.0 mM diclofenac (<b>B</b>). Effect on the response induced by 20.0 mM proline by treatments 2–5 and 0.1 mM L-Glu as a control of a negatively charged amino acid. Each value represents the mean ± SEM (n = 4–6).</p>
Full article ">Figure 6
<p>AlphaFold2-predicted structure of hPAT1 and docking poses of amino acids and NSAIDs. (<b>A</b>). AlphaFold2 structure of hPAT1 showing the proposed binding sites of glycine (in yellow) and naproxen (in green). (<b>B</b>). Docking poses of glycine (yellow) and L-proline (green) in the proposed orthosteric binding site of hPAT1. Yellow dashed lines: hydrogen bonds; green dashed lines: pi-cation interactions. (<b>C</b>–<b>G</b>). Docking poses of naproxen (<b>C</b>), ibuprofen (<b>D</b>), flurbiprofen (<b>E</b>), ketoprofen (<b>F</b>), and diclofenac (<b>G</b>) in the proposed allosteric binding site of hPAT1. Ibuprofen, flurbiprofen, and ketoprofen are racemic mixtures, and the stereoisomers with the best docking scores are depicted. Yellow dashed lines: hydrogen bonds; green dashed lines: π-cation interactions; blue dashed lines: π-π stacking; pink dashed lines: salt bridges.</p>
Full article ">
10 pages, 2847 KiB  
Article
Hydrogen-Free Plasma Nitriding Process for Fabrication of Expanded Austenite Layer on AISI 316 Stainless Steel Surface
by Mitsuhiro Hirano, Koyo Miura and Naofumi Ohtsu
Materials 2025, 18(1), 140; https://doi.org/10.3390/ma18010140 - 1 Jan 2025
Viewed by 313
Abstract
The addition of hydrogen to nitrogen facilitates the formation of nitride phases in the plasma nitriding processes of stainless steels, though it also induces the deterioration of their mechanical properties. This study presents a hydrogen-free plasma nitriding process for fabricating a nitrogen-expanded austenite [...] Read more.
The addition of hydrogen to nitrogen facilitates the formation of nitride phases in the plasma nitriding processes of stainless steels, though it also induces the deterioration of their mechanical properties. This study presents a hydrogen-free plasma nitriding process for fabricating a nitrogen-expanded austenite phase (γN) on an AISI 316 stainless steel surface. The steel substrate was nitrided in N2-Ar plasma with various gas compositions discharged by radio frequency (RF) and direct current (DC) modes. The process using the RF mode enabled the fabrication of a layer composed of a γN phase with a thickness of approximately 3 μm on the steel surface regardless of the gas composition, thereby enhancing its surface hardness. In contrast, such a layer was not observed in the DC mode, and the steel’s hardness was similar to that of the untreated surface. This difference in layer formation was attributed to the mitigation of surface etching by the Ar active species using the RF mode because of the lower bias voltage compared with that of the DC mode. This phenomenon suppresses the removal of the nitride phase formed during the process, which is a key factor contributing to nitrogen penetration. In conclusion, an N2-Ar plasma nitriding process using the RF mode is demonstrated to be a hydrogen-free process for fabricating a layer of a γN phase. Full article
(This article belongs to the Special Issue Surface Modifications and Coatings for Metallic Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of plasma nitriding apparatus with RF and DC discharge modes.</p>
Full article ">Figure 2
<p>Variations in substrate surface temperature against nitriding duration: (<b>a</b>) RF glow discharge and (<b>b</b>) DC glow discharge mode.</p>
Full article ">Figure 3
<p>XRD patterns of 316SS surface treated in N<sub>2</sub>-Ar plasmas: (<b>a</b>) RF glow discharge and (<b>b</b>) DC glow discharge mode.</p>
Full article ">Figure 4
<p>Cross-sectional SEM images of (<b>a</b>) 10% Ar and (<b>b</b>) 70% Ar surfaces treated in N<sub>2</sub>-Ar plasma generated using RF discharge and (<b>c</b>) 70% Ar surface treated by DC discharge.</p>
Full article ">Figure 5
<p>Vickers hardness of 316SS surfaces nitrided in N<sub>2</sub>-Ar plasmas generated using RF and DC glow discharge modes.</p>
Full article ">Figure 6
<p>Typical optical emission spectra of N<sub>2</sub>-Ar plasmas corresponding to 50% Ar conditions generated using RF and DC glow discharge modes.</p>
Full article ">Figure 7
<p>Emission intensities attributed to (<b>a</b>) N<sub>2</sub><sup>+</sup> (391.4 nm) and (<b>b</b>) excited Ar (811.4 nm), plotted against gas composition of Ar.</p>
Full article ">Figure 8
<p>SPM images of 316SS surfaces nitrided in N<sub>2</sub>-Ar plasmas generated by RF (<b>a</b>,<b>b</b>) and DC glow discharge modes (<b>c</b>): (<b>a</b>) 10% Ar, (<b>b</b>,<b>c</b>) 70% Ar.</p>
Full article ">
25 pages, 3570 KiB  
Review
Contemporary Trends in Pulsed Field Ablation for Cardiac Arrhythmias
by Hagai Yavin, Mark Prasad, Jonathan Gordon, Tolga Aksu and Henry D. Huang
J. Cardiovasc. Dev. Dis. 2025, 12(1), 10; https://doi.org/10.3390/jcdd12010010 - 30 Dec 2024
Viewed by 383
Abstract
Pulsed field ablation (PFA) is a catheter-based procedure that utilizes short high voltage and short-duration electrical field pulses to induce tissue injury. The last decade has yielded significant scientific progress and quickened interest in PFA as an energy modality leading to the emergence [...] Read more.
Pulsed field ablation (PFA) is a catheter-based procedure that utilizes short high voltage and short-duration electrical field pulses to induce tissue injury. The last decade has yielded significant scientific progress and quickened interest in PFA as an energy modality leading to the emergence of the clinical use of PFA technologies for the treatment of atrial fibrillation. It is generally agreed that more research is needed to improve our biophysical understanding of PFA for clinical cardiac applications as well as its potential as a potential alternative energy source to thermal ablation modalities for the treatment of other arrhythmias. In this review, we discuss the available preclinical and clinical evidence for PFA for atrial fibrillation, developments for ventricular arrhythmia (VA) ablation, and future perspectives. Full article
(This article belongs to the Special Issue Heart Rhythm Disorders: Diagnosis, Treatment, and Management)
Show Figures

Figure 1

Figure 1
<p>Conceptual figures showing an association between proximity of the electrode and relationship between strength of the electric field which may result in irreversible and reversible cellular electroporation. Theoretical differences between unipolar and bipolar configurations on biophysics of pulsed electric field delivery.</p>
Full article ">Figure 2
<p>Clinical and investigation catheter technologies for pulsed field ablation.</p>
Full article ">Figure 3
<p>(<b>A</b>) In four swine, direct ablation with PFA and RFA within the lumen of the esophagus were performed to assess the effect of PFA on esophagus tissue. <b>Left</b>, 3D anatomical map of the esophagus and RA. Red dots represent RFA while green dots, PFA. <b>Middle,</b> gross pathology demonstrates direct ablation to the esophageal lumen, interchangeably with PFA and RFA. <b>Right</b>, histological slides of PF and RF lesions show mild edema and focal superficial necrosis in PFA lesions, while RFA shows severe edema, necrosis, and hemorrhage spanning to the deep muscularis layers. (<b>B</b>) In six swine, 5.5 (1–8) PFA applications were placed on the endocardial RA, opposing the phrenic nerve. These did not result in phrenic nerve paralysis. Comparison RF ablation. <b>Left</b>, anatomical map with the course of the right phrenic nerve identified by pacing the lateral RA marked in light-blue tags. Green tags represent PFA and red represent RFA. <b>Middle</b>, gross pathology of the phrenic nerve with clear lesions at RFA sites opposed to the healthy-looking tissue at the PFA sites. <b>Right</b>, histological analysis at PFA application sites demonstrating PFA selectively affected cardiomyocytes but spared blood vessels and nervous tissue.</p>
Full article ">Figure 4
<p>(<b>A</b>) First pulsed field ablation (PFA) application from left atrium near right superior pulmonary vein induces a profound vagal response. (<b>B</b>) The red spherical tags show radiofrequency lesions given nearby the PFA lesion where acute vagal response was obtained which are indicated with blue and green spherical tags (from superior view). Although radiofrequency (RF) lesions did not induce further vagal response after PFA application, RF applications were performed to ensure long-term parasympathetic denervation. Image reproduced with permission from Sikiric et al. <span class="html-italic">J. Interv. Card Electrophysiol.</span> (2024).</p>
Full article ">Figure 5
<p>(<b>A</b>) shows penta-spline pulsed field ablation (PFA) catheter positioned at the right superior ganglionic plexus. (<b>B</b>) shows penta-spline PFA catheter position at the left superior and right inferior ganglionic plexuses.</p>
Full article ">
8 pages, 1652 KiB  
Article
Significantly Enhanced Acidic Oxygen Evolution Reaction Performance of RuO2 Nanoparticles by Introducing Oxygen Vacancy with Polytetrafluoroethylene
by Jinyang Zhang, Xinru Wang, Xinyue Zhao, Honglei Chen and Peng Jia
Polymers 2025, 17(1), 59; https://doi.org/10.3390/polym17010059 - 29 Dec 2024
Viewed by 600
Abstract
The supported RuO2 catalysts are known for their synergistic and interfacial effects, which significantly enhance both catalytic activity and stability. However, polymer-supported RuO2 catalysts have received limited attention due to challenges associated with poor conductivity. In this study, we successfully synthesized [...] Read more.
The supported RuO2 catalysts are known for their synergistic and interfacial effects, which significantly enhance both catalytic activity and stability. However, polymer-supported RuO2 catalysts have received limited attention due to challenges associated with poor conductivity. In this study, we successfully synthesized the RuO2-polytetrafluoroethylene (PTFE) catalyst via a facile annealing process. The optimized nucleation and growth strategies enable the formation of RuO2 particles (~13.4 nm) encapsulating PTFE, establishing a conductive network that effectively addresses the conductivity issue. Additionally, PTFE induces the generation of oxygen vacancies and the formation of stable RuO2/PTFE interfaces, which further enhance the acidic OER activity and the stability of RuO2. As a result, the RuO2-PTFE catalyst exhibits a low overpotential of 219 mV at 10 mA cm⁻2 in the three-electrode system, and the voltage of the RuO2-PTFE||commercial Pt/C system can keep 1.50 V for 800 h at 10 mA cm−2. This work underscores the versatility of PTFE as a substrate for fine-tuning the catalyst morphology, the crystal defect, and the stable interface outerwear. This work not only broadens the application scope of PTFE in catalyst synthesis but also provides a novel approach to the design of high-performance metallic oxide catalysts with tailored oxygen vacancy concentration and stable polymer outerwear. Full article
(This article belongs to the Special Issue Polymer-Based Smart Materials: Preparation and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>c</b>) SEM, TEM, and HRTEM images of RuO<sub>2</sub>-PTFE; (<b>d</b>–<b>f</b>) SEM, TEM, and HRTEM images of RuO<sub>2</sub>-F; (<b>g</b>) STEM image and corresponding Ru, O, and F mapping images of RuO<sub>2</sub>-PTFE.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of PTFE, RuO<sub>2</sub>-PTFE, and RuO<sub>2</sub>-F; EPR spectra of (<b>b</b>) RuO<sub>2</sub>-PTFE and (<b>c</b>) RuO<sub>2</sub>-F; high-resolution XPS spectra of (<b>d</b>) Ru 3p, (<b>e</b>) O 2p, and (<b>f</b>) F 1s for RuO<sub>2</sub>-PTFE and RuO<sub>2</sub>-F; (<b>g</b>) CV curve of RuO<sub>2</sub>-PTFE for DEMS test; DEMS spectra of (<b>h</b>) RuO<sub>2</sub>-PTFE and (<b>i</b>) RuO<sub>2</sub>-F.</p>
Full article ">Figure 3
<p>(<b>a</b>) LSV curves; (<b>b</b>) Tafel plots; (<b>c</b>) Nyquist plots; CV curves of (<b>d</b>) RuO<sub>2</sub>-PTFE and (<b>e</b>) RuO<sub>2</sub>-F; (<b>f</b>) C<sub>dl</sub> calculated by (<b>d</b>,<b>e</b>) in the three-electrode system.</p>
Full article ">Figure 4
<p>(<b>a</b>) Two-electrode performance of commercial RuO<sub>2</sub>-PTFE||commercial Pt/C catalysts; (<b>b</b>) the curves of voltage versus time for the galvanostatic stability test.</p>
Full article ">
20 pages, 8449 KiB  
Article
Recovery Analysis of Sequentially Irradiated and NBT-Stressed VDMOS Transistors
by Snežana Djorić-Veljković, Emilija Živanović, Vojkan Davidović, Sandra Veljković, Nikola Mitrović, Goran Ristić, Albena Paskaleva, Dencho Spassov and Danijel Danković
Micromachines 2025, 16(1), 27; https://doi.org/10.3390/mi16010027 - 28 Dec 2024
Viewed by 317
Abstract
This study investigates the effects of negative bias temperature (NBT) stress and irradiation on the threshold voltage (VT) of p-channel VDMOS transistors, focusing on degradation, recovery after each type of stress, and operational behavior under varying conditions. Shifts in V [...] Read more.
This study investigates the effects of negative bias temperature (NBT) stress and irradiation on the threshold voltage (VT) of p-channel VDMOS transistors, focusing on degradation, recovery after each type of stress, and operational behavior under varying conditions. Shifts in VTVT) were analyzed under different stress orders, showing distinct influence mechanisms, including defects creation and their removal and electrochemical reactions. Recovery data after each type of stress indicated ongoing electrochemical processes, influencing subsequent stress responses. Although the ΔVT is not particularly pronounced during the recovery after irradiation, changes in subthreshold characteristics indicate the changes in defect densities that affect the behavior of the components during further application. Additionally, the findings show that the ΔVT during the NBT stress after irradiation (up to certain doses and conditions) remains relatively stable, but this is the result of a balance of competing mechanisms. A subthreshold characteristic analysis provided a further insight into the degradation dynamics. A particular attention was paid to analyzing ΔVT with a focus on predicting the lifetime. In practical applications, especially under pulsed operation, prior stresses altered the device’s thermal and electrical performance. It was shown that self-heating effects were more pronounced in pre-stressed components, increasing the power dissipation and thermal instability. These insights additionally highlight the importance of understanding stress-induced degradation and recovery mechanisms for optimizing VDMOS transistor reliability in advanced electronic systems. Full article
Show Figures

Figure 1

Figure 1
<p>Subthreshold (<b>a</b>) and above-threshold (<b>b</b>) transfer characteristics of n-channel power VDMOSFETs (of two different manufacturers—M1 and M2) during long-term recovery after applied irradiation.</p>
Full article ">Figure 2
<p>Schematic representation of the steps involved in both parts of the experiment, during which the components were subjected to (<b>a</b>) NBT stress followed by irradiation and (<b>b</b>) irradiation followed by NBT stress.</p>
Full article ">Figure 3
<p>Components’ transfer characteristic shifts induced by NBT stress followed by irradiation with a gate voltage of −10 V (<b>a</b>) up to 30 Gy and (<b>b</b>) up to 120 Gy, and transfer characteristics after the spontaneous recovery that tailed both stresses.</p>
Full article ">Figure 4
<p>Components’ transfer characteristic shifts induced by irradiation with a gate voltage of −10 V (<b>a</b>) up to 30 Gy and (<b>b</b>) up to 120 Gy, followed by NBT stress, and transfer characteristics after the spontaneous recovery that tailed both stresses.</p>
Full article ">Figure 5
<p>Threshold voltage shift through NBT stress followed by irradiation and during the spontaneous recovery that tailed both phases.</p>
Full article ">Figure 6
<p>Threshold voltage shift through irradiation followed by NBT stress and during the spontaneous recovery that tailed both phases.</p>
Full article ">Figure 7
<p>Threshold voltage shift through irradiation and the spontaneous recovery of fresh and previously NBT-stressed components.</p>
Full article ">Figure 8
<p>Threshold voltage shift through NBT stress and the spontaneous recovery of fresh and previously irradiated components.</p>
Full article ">Figure 9
<p>Schematic illustration of two parts—contributions to the increase and decrease in threshold voltage shift, which are caused by the mechanisms of activation of electrochemical reactions and annealing.</p>
Full article ">Figure 10
<p>Changes in threshold voltage shift through NBT stress, after irradiation with no gate voltage applied up to four achieved doses (<b>a</b>) and lifetime prediction (<b>b</b>).</p>
Full article ">Figure 11
<p>The temperature change of two stressed groups of components, NBT-RAD and RAD-NBT, with four different drain currents.</p>
Full article ">
17 pages, 3372 KiB  
Article
An Electrical Method to Detect Both Crack Creation and Propagation in Solid Electrical Insulators
by Tara Niakan, Zarel Valdez-Nava and David Malec
Materials 2025, 18(1), 24; https://doi.org/10.3390/ma18010024 - 25 Dec 2024
Viewed by 273
Abstract
Fracto-emission is the ejection of electrons and positive ions from matter undergoing a mechanical fracture. The creation and propagation of fractures in insulating material can generate an electrical signal that can be detected using a sufficiently fast signal recorder. The theoretical equations related [...] Read more.
Fracto-emission is the ejection of electrons and positive ions from matter undergoing a mechanical fracture. The creation and propagation of fractures in insulating material can generate an electrical signal that can be detected using a sufficiently fast signal recorder. The theoretical equations related to crack creation/propagation that induce an externally electric signal are detailed for two conditions: with and without an external applied electric voltage. Results from an experiment with no externally applied voltage are presented for fibreglass-reinforced epoxy laminate samples, in which current signals ranging from 50 mA to 100 mA are measured in a time frame of 200 ns. The signal-to-noise ratio is high enough to consider that the signal that was recorded is not a measurement artifact. This method may help to identify and track a crack propagating inside dielectric materials. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>Left</b>) pristine insulator exhibiting a pre-existing crack. (<b>Right</b>) propagated crack after the 3PB test.</p>
Full article ">Figure 2
<p>(<b>Left</b>) pristine insulator with no charge (neutral). (<b>Right</b>) charge separation due to the pre-existing crack propagation.</p>
Full article ">Figure 3
<p>Solid insulator sample with two electrodes.</p>
Full article ">Figure 4
<p>Glass-reinforced epoxy laminate (PCB-FR4), 1 mm thickness, 60 mm length and 30 mm width. The copper metallization on both side has a thickness of 35 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m, with (<b>a</b>) the top side of the sample and (<b>b</b>) the bottom side.</p>
Full article ">Figure 5
<p>Sample equivalent circuit.</p>
Full article ">Figure 6
<p>Glass-reinforced epoxy laminate (PCB-FR4) (<b>a</b>) Pristine sample with metallization and (<b>b</b>) Sample placed under the 3-point bending.</p>
Full article ">Figure 7
<p>Release of a pressure wave under a mechanical constraint (3-point bending) in the insulator when a crack is created and/or propagates.</p>
Full article ">Figure 8
<p>Fracto-emission by the creation of a crack in the dielectric under a 3-point bending test.</p>
Full article ">Figure 9
<p>Charges induced by a crack creation and/or propagation. (1) Refers to charge recombination on the crack surface, (2) refers to charge recombination through the volume, and (3) refers to charge recombination across the crack.</p>
Full article ">Figure 10
<p>Evolution of the surface charge density with the creation and/or propagation of a crack with an external applied voltage. As seen before, charge recombination can occur (1) on the crack surface, (2) through the volume, or (3) across the crack. A pressure wave (4) can also occur under this condition.</p>
Full article ">Figure 11
<p>Acquisition circuit.</p>
Full article ">Figure 12
<p>Sample equivalent circuit with the acquisition circuit.</p>
Full article ">Figure 13
<p>Experimental procedure for the electrical acquisition of the crack propagation: (<b>a</b>) Sample positioned under 3-point bending linked to the electric acquisition, (<b>b</b>) sample under applied force using the 3-point bending device linked to the electric acquisition, (<b>c</b>) sample placed in the 3-point bending device with connecting cable linked to the electric acquisition, (<b>d</b>) crocodile clips attached to the sample, upper image, the top side view and bottom image, the back side, (<b>e</b>) the experimental setup with (<b>e1</b>) the 3-point bending machine, (<b>e2</b>) amplifier power supply, (<b>e3</b>) HF amplifier, (<b>e4</b>) wire between the amplifier output and the oscilloscope, (<b>e5</b>) epoxy sample, and (<b>e6</b>) AC 50 Hz high voltage for testing under an applied voltage.</p>
Full article ">Figure 14
<p>Three phases of the acquisition, (i) the first on being the initialization of the test, (ii) followed by the application of an external force and finally (iii) reaching the critical value to induced a current by propagating a crack (S1).</p>
Full article ">Figure 15
<p>Typical crack-induced current recorded during a 3-point bending (before the final mechanical breakdown) on an epoxy sample (S2).</p>
Full article ">Figure 16
<p>Typical crack-induced currents recorded during a 3-point bending, on an epoxy sample (S3). Both (<b>a</b>,<b>b</b>) peaks occurred before the final mechanical breakdown.</p>
Full article ">Figure 17
<p>Five crack-induced current signals recorded during the crack propagation using a 3-point bending (before the final mechanical breakdown) on an epoxy sample.</p>
Full article ">Figure 18
<p>Crack-induced currents recorded during a 3-point bending up to the sample final mechanical breakdown. (a) and (b) are external currents recorded during the test; (c) is the last external current recorded when the sample brakes.</p>
Full article ">
12 pages, 3116 KiB  
Article
Origin of the Temperature Dependence of Gate-Induced Drain Leakage-Assisted Erase in Three-Dimensional nand Flash Memories
by David G. Refaldi, Gerardo Malavena, Luca Chiavarone, Alessandro S. Spinelli and Christian Monzio Compagnoni
Micromachines 2024, 15(12), 1516; https://doi.org/10.3390/mi15121516 - 20 Dec 2024
Viewed by 471
Abstract
Through detailed experimental and modeling activities, this paper investigates the origin of the temperature dependence of the Erase operation in 3D nand flash arrays. First of all, experimental data collected down to the cryogenic regime on both charge-trap and floating-gate arrays are provided [...] Read more.
Through detailed experimental and modeling activities, this paper investigates the origin of the temperature dependence of the Erase operation in 3D nand flash arrays. First of all, experimental data collected down to the cryogenic regime on both charge-trap and floating-gate arrays are provided to demonstrate that the reduction in temperature makes cells harder to Erase irrespective of the nature of their storage layer. This evidence is then attributed to the weakening, with the decrease in temperature, of the gate-induced drain leakage (GIDL) current exploited to set the electrostatic potential of the body of the nand strings during Erase. Modeling results for the GIDL-assisted Erase operation, finally, allow not only to support this conclusion but also to directly correlate the change with temperature of the electrostatic potential of the string body with the change with temperature of the erased threshold-voltage of the memory cells. Full article
(This article belongs to the Section E:Engineering and Technology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic description of the test elements investigated in this work (only some of the WLs are depicted in the scheme). (<b>b</b>) Voltage waveforms applied to the array WL and SG to achieve ISPE, with grounded SL and BL (<math display="inline"><semantics> <msub> <mi>t</mi> <mi>E</mi> </msub> </semantics></math> is the duration of the Erase pulses).</p>
Full article ">Figure 2
<p>(<b>a</b>) <span class="html-italic">T</span>-induced change of cell <math display="inline"><semantics> <msub> <mi>V</mi> <mi>T</mi> </msub> </semantics></math>, with respect to 300 K. Data are reported for 50 SS and 1 MS measurements on CT-based cells. The average of the SS measurements is also highlighted. The inset shows a representative example for the <math display="inline"><semantics> <mrow> <msub> <mi>I</mi> <mrow> <mi>B</mi> <mi>L</mi> </mrow> </msub> <mo>−</mo> <msub> <mi>V</mi> <mrow> <mi>W</mi> <mi>L</mi> </mrow> </msub> </mrow> </semantics></math> trans-characteristic of the MS structure, at different <span class="html-italic">T</span> values (a log scale is used for <math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>B</mi> <mi>L</mi> </mrow> </msub> </semantics></math>). (<b>b</b>) <math display="inline"><semantics> <msub> <mi>V</mi> <mi>T</mi> </msub> </semantics></math> evolution with the ISPE pulse number at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math> K and 15 K, as obtained from MS measurements on CT-based samples (circles). ISPP results are also shown (squares). (<b>c</b>) Same as in (<b>b</b>) but with the <span class="html-italic">T</span> dependence of the Read operation subtracted from the <math display="inline"><semantics> <msub> <mi>V</mi> <mi>T</mi> </msub> </semantics></math> values at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> K. Dashed black lines are guides to the eye. All the voltages reported in this work were normalized to the same arbitrary constant.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Same as in <a href="#micromachines-15-01516-f002" class="html-fig">Figure 2</a>b,c, but for FG-based cells. (<b>c</b>,<b>d</b>) Same as in <a href="#micromachines-15-01516-f002" class="html-fig">Figure 2</a>b,c but for an average of 20 SS measurements on FG-based cells.</p>
Full article ">Figure 4
<p>Schematic band diagram along the gate stack of a CT-based memory cell during an ISPE pulse at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math> K (solid black curves) and 15 K (dashed red curves). <math display="inline"><semantics> <msub> <mi>E</mi> <mi>C</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>E</mi> <mi>V</mi> </msub> </semantics></math> are, respectively, the conduction band and valence band edge of the materials. <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math> is the Fermi level in the metal WL. For the sake of simplicity, the tunnel–dielectric stack is considered as a single-layer dielectric.</p>
Full article ">Figure 5
<p>Compact model for GIDL-assisted Erase in CT-based 3D <span class="html-small-caps">nand</span> Flash strings, adapted from [<a href="#B13-micromachines-15-01516" class="html-bibr">13</a>,<a href="#B14-micromachines-15-01516" class="html-bibr">14</a>]. The light-blue region corresponds to the part of the model reproducing the <math display="inline"><semantics> <msub> <mi>V</mi> <mi>B</mi> </msub> </semantics></math> dynamics in the presence of the GIDL current, while the yellow region corresponds to the part of the model reproducing the dynamics of charge transfer and storage in the gate stack of the cells. Note that only half of the string is considered in the model thanks to its symmetry.</p>
Full article ">Figure 6
<p><math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>G</mi> <mi>I</mi> <mi>D</mi> <mi>L</mi> </mrow> </msub> </semantics></math> vs. <math display="inline"><semantics> <msub> <mi>V</mi> <mi>B</mi> </msub> </semantics></math> characteristics assumed in the compact model simulations at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math> K and 15 K. The inset shows the factor <math display="inline"><semantics> <mi>γ</mi> </semantics></math> determining the decrease in <math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>G</mi> <mi>I</mi> <mi>D</mi> <mi>L</mi> </mrow> </msub> </semantics></math> with the reduction of <span class="html-italic">T</span> from 300 K, as obtained from experimental data collected on planar bulk MOSFETs [<a href="#B44-micromachines-15-01516" class="html-bibr">44</a>,<a href="#B45-micromachines-15-01516" class="html-bibr">45</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Simulated <math display="inline"><semantics> <msub> <mi>V</mi> <mi>B</mi> </msub> </semantics></math> transient during ISPE pulses at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math> K and 15 K. The grey and white regions correspond, respectively, to the pulse front and plateau. (<b>b</b>) Simulated <math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>B</mi> <mi>L</mi> </mrow> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>I</mi> <mrow> <mi>G</mi> <mi>I</mi> <mi>D</mi> <mi>L</mi> </mrow> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mi>I</mi> <mi>h</mi> </msub> </semantics></math> during two Erase pulses of the ISPE scheme. The grey and white regions correspond, respectively, to the pulse front and plateau.</p>
Full article ">Figure 8
<p>(<b>a</b>) Simulated <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>V</mi> <mi>T</mi> </msub> </mrow> </semantics></math> transients during ISPE at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>15</mn> </mrow> </semantics></math> K and 300 K. (<b>b</b>) Scatter plot of the vertical shift in <math display="inline"><semantics> <mrow> <mo>Δ</mo> <msub> <mi>V</mi> <mi>T</mi> </msub> </mrow> </semantics></math> transient with <span class="html-italic">T</span> and its corresponding <math display="inline"><semantics> <mrow> <mi>δ</mi> <msub> <mi>V</mi> <mi>B</mi> </msub> </mrow> </semantics></math> for different <span class="html-italic">T</span> values between 15 K and 300 K. The black dashed line is a guide to the eye, highlighting a one-to-one correlation.</p>
Full article ">
15 pages, 8086 KiB  
Article
Analysis of Measurements of the Magnetic Flux Density in Steel Blocks of the Compact Muon Solenoid Magnet Yoke with Solenoid Coil Fast Discharges
by Vyacheslav Klyukhin, Benoit Curé, Andrea Gaddi, Antoine Kehrli, Maciej Ostrega and Xavier Pons
Symmetry 2024, 16(12), 1689; https://doi.org/10.3390/sym16121689 - 19 Dec 2024
Viewed by 542
Abstract
The general-purpose Compact Muon Solenoid (CMS) detector at the Large Hadron Collider (LHC) at CERN is used to study the production of new particles in proton–proton collisions at an LHC center of mass energy of 13.6 TeV. The detector includes a magnet based [...] Read more.
The general-purpose Compact Muon Solenoid (CMS) detector at the Large Hadron Collider (LHC) at CERN is used to study the production of new particles in proton–proton collisions at an LHC center of mass energy of 13.6 TeV. The detector includes a magnet based on a 6 m diameter superconducting solenoid coil operating at a current of 18.164 kA. This current creates a central magnetic flux density of 3.8 T that allows for the high-precision measurement of the momenta of the produced charged particles using tracking and muon subdetectors. The CMS magnet contains a 10,000 ton flux-return yoke of dodecagonal shape made from the assembly of construction steel blocks distributed in several layers. These steel blocks are magnetized with the solenoid returned magnetic flux and wrap the muons escaping the hadronic calorimeters of total absorption. To reconstruct the muon trajectories, and thus to measure the muon momenta, the drift tube and cathode strip chambers are located between the layers of the steel blocks. To describe the distribution of the magnetic flux in the magnet yoke layers, a three-dimensional computer model of the CMS magnet is used. To validate the calculations, special measurements are performed, with the flux loops wound in 22 cross-sections of the flux-return yoke blocks. The measured voltages induced in the flux loops during the CMS magnet ramp-ups and -downs, as well as during the superconducting coil fast discharges, are integrated over time to obtain the initial magnetic flux densities in the flux loop cross-sections. The measurements obtained during the seven standard ramp-downs of the magnet were analyzed in 2018. From that time, three fast discharges occurred during the standard ramp-downs of the magnet. This allows us to single out the contributions of the eddy currents, induced in steel, to the flux loop voltages registered during the fast discharges of the coil. Accounting for these contributions to the flux loop measurements during intentionally triggered fast discharges in 2006 allows us to perform the validation of the CMS magnet computer model with better precision. The technique for the flux loop measurements and the obtained results are presented and discussed. The method for measuring magnetic flux density in steel blocks described in this study is innovative. The experience of 3D modeling and measuring the magnetic field in steel blocks of the magnet yoke, as part of a muon detector system, has good prospects for use in the construction and operation of particle detectors for the Future Circular Electron–Positron Collider and the Circular Electron–Positron Collider. Full article
(This article belongs to the Section Physics)
Show Figures

Figure 1

Figure 1
<p>Modeled distribution of the magnetic flux density <span class="html-italic">B</span> in Tesla in the vertical <span class="html-italic">YZ</span>-plane of the flux loop location area. Sixteen flux loops are installed in the 30° azimuthal sector at 270° of the CMS magnet barrel flux-return yoke on four layers (TC, L1, L2, L3) of the central barrel wheel YB0, and on three layers (L1, L2, L3) of the barrel wheels YB−1 and YB−2 at negative <span class="html-italic">Z</span>-coordinates, shown in meters on the <span class="html-italic">Z</span>-axis. Six flux loops are wound at the distances 1, 2, and 3 along the 18° azimuthal sector at 270° of the CMS endcap thick disks YE−1 and YE−2 [<a href="#B6-symmetry-16-01689" class="html-bibr">6</a>]. The black lines indicate the <span class="html-italic">Y</span>- or <span class="html-italic">Z</span>-positions of the loop cross-sections. The small black squares indicate the projections to the <span class="html-italic">YZ</span>-plane of the 3D B-sensor locations installed on the surfaces of the yoke steel blocks.</p>
Full article ">Figure 2
<p>Measured magnetic current variations during the fast discharges which occurred from the CMS magnet currents of 9.5, 12.5, 15, 15.221, 17.55, 18.164, and 19.14 kA. These currents create initial central magnetic flux densities of <span class="html-italic">B</span><sub>0</sub> of 2.02, 2.64, 3.16, 3.20, 3.68, 3.81, and 4.01 T, respectively, in the CMS superconducting coil.</p>
Full article ">Figure 3
<p>DAQ block diagram of the flux loop signal processing. The measuring system contains six custom multiplexers assembled in the patch boxes to read out the voltages induced in the flux loops. Through the custom-made master PLC interface circuit, the multiplexers communicate to Siemens S7-1500 PLC drives to propagate the signals to the WinCC OA software that stores the voltage values in the database.</p>
Full article ">Figure 4
<p>(<b>a</b>) One of six readout boxes with a custom multiplexer, a 24 V DC power supply, and a 19-pin Burndy connector (<b>b</b>) wiring from three to six flux loops to each multiplexer.</p>
Full article ">Figure 5
<p>(<b>a</b>) Voltages measured in the flux loop P on the L2 layer of the external barrel wheel YB−2 during all seven fast discharges performed with different initial magnet currents. The orange short-dashed line cuts the contribution of the eddy currents to the signal at the beginning of the magnet fast discharge performed from the current of 19.14 kA. This contribution has been estimated [<a href="#B6-symmetry-16-01689" class="html-bibr">6</a>] at the level of 4.6% into the integrated magnetic flux. (<b>b</b>) Voltages measured in the middle flux loop 2 of the endcap disk YE−2 during five fast discharges performed from different magnet currents.</p>
Full article ">Figure 6
<p>Three fast discharges occasionally occurred from the currents of 9.5 (on 30 November 2017), 15.221 (on 8 August 2023), and 18.164 (on 22 March 2023) kA during the standard CMS magnet ramp-downs with a rate of 1–1.5 A/s from the operational magnet current of 18.164 kA.</p>
Full article ">Figure 7
<p>The eddy current contributions to 20 flux loops vs. initial currents of the fast discharges.</p>
Full article ">Figure 8
<p>Initial axial <span class="html-italic">B<sub>z</sub></span> (negative) magnetic flux density or vertical <span class="html-italic">B<sub>y</sub></span> (positive) magnetic flux density components calculated in all 22 flux loop cross-sections at the seven CMS coil currents.</p>
Full article ">Figure 9
<p>Initial axial magnetic flux density or vertical magnetic flux density components measured in all 22 flux loop cross-sections at the seven CMS coil currents.</p>
Full article ">Figure 10
<p>Comparison of measured and calculated axial magnetic flux density or vertical magnetic flux density components in all 22 flux loop cross-sections at the seven CMS coil currents.</p>
Full article ">Figure 11
<p>Averaged comparison of measured and calculated axial magnetic flux density, <span class="html-italic">B<sub>z</sub></span> (in the 16 barrel flux loops), or vertical magnetic flux density, <span class="html-italic">B<sub>y</sub></span> (in the 6 endcap flux loops), as well as of both components in all 22 flux loop cross-sections (yoke), versus the CMS coil currents of 9.5, 12.5, 15, 15.221, 17.55, 18.164, and 19.14 kA. The currents of 9.5, 15.221, and 18.164 kA are the reference currents, where the eddy current contributions are obtained directly from the data.</p>
Full article ">Figure 12
<p>Comparison of measured and calculated axial magnetic flux density, <span class="html-italic">B<sub>z</sub></span>, or vertical magnetic flux density, <span class="html-italic">B<sub>y</sub></span>, in each of the 22 flux loop cross-sections performed in different sets of measurements: new measurements at 9.5, 15.221, and 18.164 kA; old measurements at 12.5, 15, 17.55, and 19.14 kA; and both new and old (all) measurements. The comparisons of the axial magnetic flux density measured with 3D B-sensors [<a href="#B6-symmetry-16-01689" class="html-bibr">6</a>] and calculated with the CMS magnet model [<a href="#B7-symmetry-16-01689" class="html-bibr">7</a>] are also displayed.</p>
Full article ">
14 pages, 2036 KiB  
Article
New Label-Free DNA Nanosensor Based on Top-Gated Metal–Ferroelectric–Metal Graphene Nanoribbon on Insulator Field-Effect Transistor: A Quantum Simulation Study
by Khalil Tamersit, Abdellah Kouzou, José Rodriguez and Mohamed Abdelrahem
Nanomaterials 2024, 14(24), 2038; https://doi.org/10.3390/nano14242038 - 19 Dec 2024
Viewed by 402
Abstract
In this paper, a new label-free DNA nanosensor based on a top-gated (TG) metal–ferroelectric–metal (MFM) graphene nanoribbon field-effect transistor (TG-MFM GNRFET) is proposed through a simulation approach. The DNA sensing principle is founded on the dielectric modulation concept. The computational method employed to [...] Read more.
In this paper, a new label-free DNA nanosensor based on a top-gated (TG) metal–ferroelectric–metal (MFM) graphene nanoribbon field-effect transistor (TG-MFM GNRFET) is proposed through a simulation approach. The DNA sensing principle is founded on the dielectric modulation concept. The computational method employed to evaluate the proposed nanobiosensor relies on the coupled solutions of a rigorous quantum simulation with the Landau–Khalatnikov equation, considering ballistic transport conditions. The investigation analyzes the effects of DNA molecules on nanodevice behavior, encompassing potential distribution, ferroelectric-induced gate voltage amplification, transfer characteristics, subthreshold swing, and current ratio. It has been observed that the feature of ferroelectric-induced gate voltage amplification using the integrated MFM structure can significantly enhance the biosensor’s sensitivity to DNA molecules, whether in terms of threshold voltage shift or drain current variation. Additionally, we propose the current ratio as a sensing metric due to its ability to consider all DNA-induced modulations of electrical parameters, specifically the increase in on-state current and the decrease in off-state current and subthreshold swing. The obtained results indicate that the proposed negative-capacitance GNRFET-based DNA nanosensor could be considered an intriguing option for advanced point-of-care testing. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Three-dimensional structure of the label-free DNA sensor based on TG-MFM GNRFET. (<b>b</b>) DNA detection based on the dielectric modulation concept. (<b>c</b>) Lengthwise cut view of the proposed nanoscale biosensor.</p>
Full article ">Figure 2
<p>(<b>a</b>) Flowchart of the computational method used. (<b>b</b>) Drain current values from the literature and our simulator and the P–E proprieties from L–K theory and reported experiment data for the ferroelectric hafnium zirconium oxide.</p>
Full article ">Figure 3
<p>Two-dimensional electron potential distribution at V<sub>DS</sub> = 0.3 V and V<sub>GS</sub> = 0.1 V for baseline TG GNRFET-based biosensor (<b>top figures</b>) and TG-MFM GNRFET-based biosensor (<b>bottom figures).</b> (<b>a</b>,<b>c</b>) Empty open cavity. (<b>b</b>,<b>d</b>) Cavity filled with DNA molecules.</p>
Full article ">Figure 4
<p>V<sub>G-INT</sub> versus V<sub>G-EXT</sub> for the proposed biosensor with different ferroelectric thicknesses considering (<b>a</b>) an empty and (<b>b</b>) filled sensing cavity.</p>
Full article ">Figure 5
<p>The I<sub>DS</sub>-V<sub>GS</sub> characteristics for (<b>a</b>) the baseline and (<b>b</b>) the proposed biosensor considering different DNA dielectric constants.</p>
Full article ">Figure 6
<p>Subthreshold swing (<b>a</b>) and current ratio (<b>b</b>) as functions of DNA dielectric constant for the TG-MFM GNRFET-based DNA sensor.</p>
Full article ">
17 pages, 10949 KiB  
Article
Research on the Detection Method for Feeding Metallic Foreign Objects in Coal Mine Crushers Based on Reflective Pulsed Eddy Current Testing
by Benchang Meng, Zezheng Zhuang, Jiahao Ma and Sihai Zhao
Appl. Sci. 2024, 14(24), 11704; https://doi.org/10.3390/app142411704 - 15 Dec 2024
Viewed by 624
Abstract
In response to the difficulties and poor timeliness in detecting feeding metallic foreign objects during high-yield continuous crushing operations in coal mines, this paper proposes a new method for detecting metallic foreign objects, combining pulsed eddy current testing with the Truncated Region Eigenfunction [...] Read more.
In response to the difficulties and poor timeliness in detecting feeding metallic foreign objects during high-yield continuous crushing operations in coal mines, this paper proposes a new method for detecting metallic foreign objects, combining pulsed eddy current testing with the Truncated Region Eigenfunction Expansion (TREE) method. This method is suitable for the harsh working conditions in coal mine crushing stations, which include high dust, strong vibration, strong electromagnetic interference, and low temperatures in winter. A model of the eddy current field of feeding metallic foreign objects in the truncated region is established using a coaxial excitation and receiving coil with a Hall sensor. The full-cycle time-domain analytical solution for the induced voltage and magnetic induction intensity of the reflective field under practical square wave signals is obtained. Simulation and experimental results show that the effective time range, peak value, and time to peak of the received voltage and magnetic induction signals can be used to classify and identify the size, thickness, conductivity, and magnetic permeability of feeding metallic foreign objects. Experimental results meet the actual needs for removing feeding metallic foreign objects in coal mine sites. This provides core technical support for the establishment of a predictive fault diagnosis system for crushing equipment. Full article
Show Figures

Figure 1

Figure 1
<p>Structure diagram of the open-pit coal mine crushing station (1—Mining Truck, 2—Ore Receiving Hopper, 3—Plate Feeder, 4—Protective Steel Structure, 5—Electrical Control Room, 6—Detection Probes Array, 7—Dual-roll Screening Crusher, and 8—Belt Conveyor).</p>
Full article ">Figure 2
<p>Structure diagram of the dual-roll screening crusher (1—Wear Plates for Front and Side Walls, 2—Crusher Tooth Rolls, 3—Drive Motor, 4—Hydraulic Coupling, 5—Reducer, and 6—Coupling).</p>
Full article ">Figure 3
<p>Side view of the truncated region of (<b>a</b>) the single-turn coil, and (<b>b</b>) the rectangular cross-section coaxial excitation and receiving coils with Hall sensors.</p>
Full article ">Figure 4
<p>Typical PEC signals with non-ferromagnetic metals; (<b>a</b>) receiving coil voltage signals; (<b>b</b>) magnetic induction signals of Hall sensor.</p>
Full article ">Figure 5
<p>Typical PEC signals with ferromagnetic metals; (<b>a</b>) receiving coil voltage signals; (<b>b</b>) magnetic induction signals of Hall sensor.</p>
Full article ">Figure 6
<p>Single-probe testing experiment; (<b>a</b>) experimental platform; (<b>b</b>) block diagram of the system.</p>
Full article ">Figure 7
<p>Detailed view of single-probe and samples; (<b>a</b>) bottom view of the single-probe; (<b>b</b>) seven test samples for experiment.</p>
Full article ">Figure 8
<p>PEC differential signals of alloy steel 42CrMo with different thicknesses; (<b>a</b>) receiving coil differential voltage signals; (<b>b</b>) magnetic induction differential signals of Hall sensor.</p>
Full article ">Figure 9
<p>Relationship between key characteristic quantities of PEC differential signals and the thicknesses of alloy steel 42CrMo; (<b>a</b>) peak voltage and its corresponding time to peak; (<b>b</b>) peak magnetic inductance and its corresponding time to peak.</p>
Full article ">Figure 10
<p>Three-dimensional surface plots between key characteristics of pulsed eddy current differential voltage signals and the conductivity and thickness of non-ferromagnetic metals; (<b>a</b>) peak voltage; (<b>b</b>) time to peak.</p>
Full article ">Figure 11
<p>Three-dimensional surface plots between key characteristics of pulsed eddy current differential magnetic inductance signals and the conductivity and thickness of non-ferromagnetic metals; (<b>a</b>) peak magnetic inductance; (<b>b</b>) time to peak.</p>
Full article ">Figure 12
<p>Field experiment platform with the multi-probe array.</p>
Full article ">Figure 13
<p>Dual <span class="html-italic">Y</span>-axis plot of PEC differential signals and time for the effective detection interval in the field experiment.</p>
Full article ">
Back to TopTop