[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,123)

Search Parameters:
Keywords = nanowires

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 4686 KiB  
Article
Comparison of Aging Performances and Mechanisms: Super-Durable Fire-Resistant “Xuan Paper” Versus Chinese Traditional Xuan Paper
by Li-Ying Dong, Ying-Jie Zhu, Jin Wu and Han-Ping Yu
Molecules 2025, 30(2), 263; https://doi.org/10.3390/molecules30020263 - 10 Jan 2025
Viewed by 274
Abstract
Paper is a thin nonwoven material made from cellulose fibers as the main raw material together with some additives. Paper is highly flammable, leading to the destruction of countless precious ancient books, documents, and art works in fire disasters. In recent years, researchers [...] Read more.
Paper is a thin nonwoven material made from cellulose fibers as the main raw material together with some additives. Paper is highly flammable, leading to the destruction of countless precious ancient books, documents, and art works in fire disasters. In recent years, researchers have made a lot of efforts in order to obtain more durable and fire-retardant paper. Owing to the successful synthesis of ultralong hydroxyapatite (HAP) nanowires as a new kind of inorganic nanofiber material, it becomes possible to develop a new kind of super-durable and fire-resistant paper. Recently, the authors’ research group prepared a new kind of fire-resistant “Xuan paper” consisting of ultralong HAP nanowires. In this article, the super-durable fire-resistant “Xuan paper” based on ultralong HAP nanowires and the traditional Xuan paper based on cellulose fibers were evaluated by the accelerated aging method for 1200 days at 105 °C in air, which is the equivalent of 10,000 years of natural aging in the ambient environment. The aging mechanism of the traditional Xuan paper was further investigated by studying the fiber length/width and their distributions, morphology, infrared spectroscopy, thermogravimetric analysis, H–nuclear magnetic resonance spectra, and C–nuclear magnetic resonance spectra of cellulose fibers before and after the accelerated aging. The durability, properties, and mechanism of the fire-resistant “Xuan paper” based on ultralong HAP nanowires during the accelerated aging were studied. The experiments reveal the reasons for the deteriorated properties and reduced durability by aging of the traditional Xuan paper based on cellulose fibers, and the mechanism for the super-durability and excellent performances of the fire-resistant “Xuan paper” based on ultralong HAP nanowires during the accelerated aging process. Full article
(This article belongs to the Section Nanochemistry)
10 pages, 3918 KiB  
Article
Design and Fabrication of Ultrathin Metallic Phase Shifters for Visible and Near-Infrared Wavelengths
by Qing Guo, Jinkui Chu, Chuanlong Guan, Chuxiao Zhang and Ran Zhang
Micromachines 2025, 16(1), 74; https://doi.org/10.3390/mi16010074 - 10 Jan 2025
Viewed by 348
Abstract
The polarization state of light is critical for biological imaging, acousto-optics, bio-navigation, and many other optical applications. Phase shifters are extensively researched for their applications in optics. The size of optical elements with phase delay that are made from natural birefringent materials is [...] Read more.
The polarization state of light is critical for biological imaging, acousto-optics, bio-navigation, and many other optical applications. Phase shifters are extensively researched for their applications in optics. The size of optical elements with phase delay that are made from natural birefringent materials is limited; however, fabricating waveplates from dielectric metamaterials is very complex and expensive. Here, we present an ultrathin (14 nm) metallic phase shifter developed using nanoimprinting technology and the oxygen plasma ashing technique for visible and near-infrared wavelengths. The fabrication process can produce desirable metallic phase shifters with high efficiency, large area, and low cost. We demonstrate through a numerical simulation and experiment that the metallic phase shifter exhibits phase delay performance. Our results highlight the simplicity of the fabrication process for a metallic phase shifter with phase delay performance and offer important opportunities for creating high-efficiency, ultrathin polarizing elements, which can be used in miniaturized devices, such as integrated circuits. Full article
(This article belongs to the Special Issue Nanostructured Optoelectronic and Nanophotonic Devices)
Show Figures

Figure 1

Figure 1
<p>Structural schematics of double-layer and single-layer aluminum nanowire gratings.</p>
Full article ">Figure 2
<p>Transmittance and phase of double-layer and single-layer aluminum nanowire gratings.</p>
Full article ">Figure 3
<p>Effect of thickness, width, and period on transmittance and phase. (<b>a</b>) Effect of thickness on transmittance of <span class="html-italic">x</span>-polarized light. (<b>b</b>) Effect of thickness on transmittance of <span class="html-italic">y</span>-polarized light. (<b>c</b>) Effect of thickness on phase. (<b>d</b>) Effect of width on transmittance of <span class="html-italic">x</span>-polarized light. (<b>e</b>) Effect of width on transmittance of <span class="html-italic">y</span>-polarized light. (<b>f</b>) Effect of width on phase. (<b>g</b>) Effect of period on transmittance of <span class="html-italic">x</span>-polarized light. (<b>h</b>) Effect of period on transmittance of <span class="html-italic">y</span>-polarized light. (<b>i</b>) Effect of period on phase.</p>
Full article ">Figure 4
<p>Near-field distribution. (<b>a</b>) Incident polarization is along the length of the nanowire gratings. (<b>b</b>) Incident polarization is along the width of the nanowire gratings. The wavelength used in the simulation is 427 nm.</p>
Full article ">Figure 5
<p>Schematic diagram of fabrication process: (<b>a</b>) preparation of flexible IPS by NIL; (<b>b</b>) transfer patterns of IPS to UV-curable resist layer using UV-NIL; (<b>c</b>) aluminum thermal evaporation process; (<b>d</b>) oxygen plasma ashing process; (<b>e</b>) desirable ultrathin metallic phase shifter.</p>
Full article ">Figure 6
<p>Manufacturing process of ultrathin metallic phase shifter: (<b>a</b>) SEM surface view of flexible IPS; (<b>b</b>) SEM cross-sectional image and (<b>c</b>) SEM surface view of feature patterns of UV-curable resist layer; (<b>d</b>) SEM cross-sectional image of double-layer aluminum nanowire gratings; (<b>e</b>) AFM surface view and (<b>f</b>) AFM cross-sectional image of double-layer aluminum nanowire gratings; (<b>g</b>) SEM cross-sectional image and (<b>h</b>) SEM surface view of ultrathin metallic phase shifter.</p>
Full article ">Figure 7
<p>Test results of transmittance and phase of ultrathin metallic phase shifter: test system schemes of (<b>a</b>) transmittance and (<b>b</b>) phase; (<b>c</b>) test results of transmittance and phase difference.</p>
Full article ">
11 pages, 2505 KiB  
Article
Enhanced Photocatalytic Oxidative Coupling of Methane over Metal-Loaded TiO2 Nanowires
by Shuang Song, Jiongcan Xiang, Hui Kang and Fengming Yang
Molecules 2025, 30(2), 206; https://doi.org/10.3390/molecules30020206 - 7 Jan 2025
Viewed by 308
Abstract
The photocatalytic oxidative coupling of methane (OCM) on metal-loaded one-dimensional TiO2 nanowires (TiO2 NWs) was performed. With metal loading, the electric and optical properties of TiO2 NWs were adjusted, contributing to the improvement of the activity and selectivity of the [...] Read more.
The photocatalytic oxidative coupling of methane (OCM) on metal-loaded one-dimensional TiO2 nanowires (TiO2 NWs) was performed. With metal loading, the electric and optical properties of TiO2 NWs were adjusted, contributing to the improvement of the activity and selectivity of the OCM reaction. In the photocatalytic OCM reaction, the 1.0 Au/TiO2 NW catalyst exhibits an outstanding C2H6 production rate (4901 μmol g−1 h−1) and selectivity (70%), alongside the minor production of C3H8 and C2H4, achieving a total C2–C3 hydrocarbon selectivity of 75%. In contrast, catalysts loaded with Ag, Pd, and Pt show significantly lower activity, with Pt/TiO2 NWs producing only CO2, indicating a propensity for the deep oxidation of methane. The O2-TPD analyses reveal that Au facilitates mild O2 adsorption and activation, whereas Pt triggers excessive oxidation. Spectroscopic and kinetic studies demonstrate that Au loading not only enhances the separation efficiency of photogenerated electron–hole pairs, but also promotes the generation of active oxygen species in moderate amounts, which facilitates the formation of methyl radicals and their coupling into C2H6 while suppressing over-oxidation to CO2. This work provides novel insights and design strategies for developing efficient photocatalysts. Full article
(This article belongs to the Special Issue Nanomaterials for Energy Storage and Conversion)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) HRTEM image of TiO<sub>2</sub> NWs. (<b>b</b>) TEM image of 1.0 Au/TiO<sub>2</sub> NWs. (<b>c</b>) The distribution of Au particle size on the surface of TiO<sub>2</sub> NWs. (<b>d</b>) HADDF-STEM image of 1.0 Au/TiO<sub>2</sub> NWs and the corresponding elemental maps for Ti, O, and Au (<b>e</b>–<b>g</b>) (scale bar: 10 nm). The inset of (<b>a</b>) is selected area electron diffraction (SAED). The inset of (<b>b</b>) is HRTEM of 1.0 Au/TiO<sub>2</sub> NWs.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of TiO<sub>2</sub> NWs and 1.0 Au/TiO<sub>2</sub> NWs. (<b>b</b>) XPS spectra of 1.0 Au/TiO<sub>2</sub> NWs for the Au 4f spectrum. (<b>c</b>) UV-vis diffuse absorption spectrum of TiO<sub>2</sub> NWs and 1.0 Au/TiO<sub>2</sub> NWs. (<b>d</b>) N<sub>2</sub> adsorption–desorption isotherms of TiO<sub>2</sub> NWs and 1.0 Au/TiO<sub>2</sub> NWs.</p>
Full article ">Figure 3
<p>Photocatalytic OCM performance. (<b>a</b>) Photocatalytic activity of OCM over TiO<sub>2</sub> NWs loaded with different cocatalysts. (<b>b</b>) Photocatalytic activity of OCM over TiO<sub>2</sub> NWs loaded with different amounts of Au. (<b>c</b>) The contrast experiments performed under dark, visible and UV-visible conditions, respectively. (<b>d</b>) The contrast experiment performed under different temperature without light. Reaction conditions in (<b>a</b>–<b>d</b>): 20 mg photocatalyst, 69 mL min<sup>−1</sup> of CH<sub>4</sub> + 1 mL min<sup>−1</sup> of air (20 vol.% O<sub>2</sub>/N<sub>2</sub>).</p>
Full article ">Figure 4
<p>(<b>a</b>) Mass spectrum (<span class="html-italic">m</span>/<span class="html-italic">z</span> = 28–32) of gas products after 30 min light irradiation using 20 mg 1.0 Au/TiO<sub>2</sub> NW catalyst and 92.56 KPa <sup>12</sup>CH<sub>4</sub>. (<b>b</b>) Isotope-labelled mass spectrum (<span class="html-italic">m</span>/<span class="html-italic">z</span> = 28–32) of gas products under the same conditions using 92.56 KPa <sup>13</sup>CH<sub>4</sub> in photocatalytic oxidative coupling of methane reaction. (<b>c</b>) O<sub>2</sub>-TPD spectra of TiO<sub>2</sub> NWs, 1.0 Au/TiO<sub>2</sub> NWs, and 1.0 Pt/TiO<sub>2</sub> NWs. (<b>d</b>) PL decay spectra of TiO<sub>2</sub> NWs, 1.0 Au/TiO<sub>2</sub> NWs, and 1.0 Pt/TiO<sub>2</sub> NWs.</p>
Full article ">Figure 5
<p>(<b>a</b>) In situ DRIFT spectroscopy characterization of 1.0 Au/TiO<sub>2</sub> NWs. (<b>b</b>) In situ DRIFT spectroscopy characterization of 1.0 Pt/TiO<sub>2</sub> NWs (in situ DRIFT spectra were test in reactant gas (CH<sub>4</sub>/air ratio, 69/1) under different light irradiation times).</p>
Full article ">
12 pages, 3591 KiB  
Article
Multilayer Graphene Stacked with Silver Nanowire Networks for Transparent Conductor
by Jinsung Kwak
Materials 2025, 18(1), 208; https://doi.org/10.3390/ma18010208 - 6 Jan 2025
Viewed by 335
Abstract
A mechanically robust flexible transparent conductor with high thermal and chemical stability was fabricated from welded silver nanowire networks (w-Ag-NWs) sandwiched between multilayer graphene (MLG) and polyimide (PI) films. By modifying the gas flow dynamics and surface chemistry of the Cu surface during [...] Read more.
A mechanically robust flexible transparent conductor with high thermal and chemical stability was fabricated from welded silver nanowire networks (w-Ag-NWs) sandwiched between multilayer graphene (MLG) and polyimide (PI) films. By modifying the gas flow dynamics and surface chemistry of the Cu surface during graphene growth, a highly crystalline and uniform MLG film was obtained on the Cu foil, which was then directly coated on the Ag-NW networks to serve as a barrier material. It was found that the highly crystalline layers in the MLG film compensate for structural defects, thus forming a perfect barrier film to shield Ag NWs from oxidation and sulfurization. MLG/w-Ag-NW composites were then embedded into the surface of a transparent and colorless PI thin film by spin-coating. This allowed the MLG/w-Ag-NW/PI composite to retain its original structural integrity due to the intrinsic physical and chemical properties of PI, which also served effectively as a binder. In view of its unique sandwich structure and the chemical welding of the Ag NWs, the flexible substrate-cum-electrode had an average sheet resistance of ≈34 Ω/sq and a transmittance of ≈91% in the visible range, and also showed excellent stability against high-temperature annealing and sulfurization. Full article
(This article belongs to the Section Advanced Nanomaterials and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) A sequential schematic diagram of fabricating MLG/Ag-NW/PI composites. First, using LPCVD in our oxidation protocols, a large-area MLG layer was grown on a Cu foil. Then, Ag-NW solution was spin-coated on an MLG/Cu foil. Followed by chemical welding of Ag-NW networks on an MLG/Cu foil, transparent soluble PI solution was spin-coated and cured for structural integrity. Finally, Cu was etched away and the MLG/w-Ag-NW/PI composite was obtained. (<b>b</b>) A photo of MLG/w-Ag-NW/PI held by tweezers. (<b>c</b>–<b>e</b>) Representative SEM images of MLG/w-Ag-NW/PI composite. Purple arrow in (<b>d</b>) indicates typical torn MLG and green arrow in (<b>e</b>) indicates a wrinkle of MLG right above Ag-NWs. (<b>f</b>) Representative Raman spectrum obtained from various locations in an MLG layer.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) Schematic illustrations of layer-controlled CVD graphene synthesis by changing the position of a Cu foil loaded in the interior of a quartz tube (upper panels) and corresponding optical microscope images of as-synthesized graphene layers transferred onto a Si/SiO<sub>2</sub> substrate (lower panels). (<b>c</b>) A schematic illustration of an air-oxidized Cu foil loaded at the bottom of a quartz tube with MLG transferred onto a Si/SiO<sub>2</sub> substrate observed using an optical microscope. (<b>d</b>) MLG coverage data as a function of Cu foil oxidation method. (<b>e</b>) Data on graphene coverage on back side of Cu foil as a function of Cu foil oxidation method. (<b>f</b>) MLG coverage data as a function of growth time.</p>
Full article ">Figure 3
<p>(<b>a</b>) Sheet resistance and transmittance at 550 nm of Gr/PI composites according to the MLG portion. (<b>b</b>) Sheet resistance and transmittance at 550 nm of MLG/Ag-NW/PI films depending on the number of Ag coating cycles, in which MLG coverage is ~98%. (<b>c</b>) Comparison of sheet resistance values before and after chemical welding of Ag-NWs on the Ag-NW/MLG/Cu foil. (<b>d</b>) Comparison of optoelectronic properties (<span class="html-italic">R</span><sub>sh</sub> and transmittance at 550 nm) between our MLG/w-AgNW/PI composite and previously reported results. The performance of Ag-NW [<a href="#B40-materials-18-00208" class="html-bibr">40</a>], graphene [<a href="#B41-materials-18-00208" class="html-bibr">41</a>], SWCNT [<a href="#B42-materials-18-00208" class="html-bibr">42</a>], and SWCNT/graphene [<a href="#B43-materials-18-00208" class="html-bibr">43</a>] composites is shown for comparison. (<b>e</b>,<b>f</b>) Mechanical bending properties of TCEs upon continuous bending cycles and bending radii, respectively.</p>
Full article ">Figure 4
<p>Test for environmental stability of our MLG/w-Ag-NW/PI composite in air (<b>a</b>) at ~25 °C and (<b>b</b>) ~100 °C and (<b>c</b>) in an aqueous ammonium persulfate solution (0.1 M), in which a change in resistance is observed.</p>
Full article ">
19 pages, 1864 KiB  
Article
An FPGA-Based SiNW-FET Biosensing System for Real-Time Viral Detection: Hardware Amplification and 1D CNN for Adaptive Noise Reduction
by Ahmed Hadded, Mossaad Ben Ayed and Shaya A. Alshaya
Sensors 2025, 25(1), 236; https://doi.org/10.3390/s25010236 - 3 Jan 2025
Viewed by 421
Abstract
Impedance-based biosensing has emerged as a critical technology for high-sensitivity biomolecular detection, yet traditional approaches often rely on bulky, costly impedance analyzers, limiting their portability and usability in point-of-care applications. Addressing these limitations, this paper proposes an advanced biosensing system integrating a Silicon [...] Read more.
Impedance-based biosensing has emerged as a critical technology for high-sensitivity biomolecular detection, yet traditional approaches often rely on bulky, costly impedance analyzers, limiting their portability and usability in point-of-care applications. Addressing these limitations, this paper proposes an advanced biosensing system integrating a Silicon Nanowire Field-Effect Transistor (SiNW-FET) biosensor with a high-gain amplification circuit and a 1D Convolutional Neural Network (CNN) implemented on FPGA hardware. This attempt combines SiNW-FET biosensing technology with FPGA-implemented deep learning noise reduction, creating a compact system capable of real-time viral detection with minimal computational latency. The integration of a 1D CNN model on FPGA hardware for adaptive, non-linear noise filtering sets this design apart from conventional filtering approaches by achieving high accuracy and low power consumption in a portable format. This integration of SiNW-FET with FPGA-based CNN noise reduction offers a unique approach, as prior noise reduction techniques for biosensors typically rely on linear filtering or digital smoothing, which lack adaptive capabilities for complex, non-linear noise patterns. By introducing the 1D CNN on FPGA, this architecture enables real-time, high-fidelity noise reduction, preserving critical signal characteristics without compromising processing speed. Notably, the findings presented in this work are based exclusively on comprehensive simulations using COMSOL and MATLAB, as no physical prototypes or biomarker detection experiments were conducted. The SiNW-FET biosensor, functionalized with antibodies specific to viral antigens, detects impedance shifts caused by antibody–antigen interactions, providing a highly sensitive platform for viral detection. A high-gain folded-cascade amplifier enhances the Signal-to-Noise Ratio (SNR) to approximately 70 dB, verified through COMSOL and MATLAB simulations. Additionally, a 1D CNN model is employed for adaptive noise reduction, filtering out non-linear noise patterns and achieving an approximate 75% noise reduction across a broad frequency range. The CNN model, implemented on an Altera DE2 FPGA, enables high-throughput, low-latency signal processing, making the system viable for real-time applications. Performance evaluations confirmed the proposed system’s capability to enhance the SNR significantly while maintaining a compact and energy-efficient design suitable for portable diagnostics. This integrated architecture thus provides a powerful solution for high-precision, real-time viral detection, and continuous health monitoring, advancing the role of biosensors in accessible point-of-care diagnostics. Full article
(This article belongs to the Special Issue Advanced Sensor Technologies for Biomedical-Information Processing)
Show Figures

Figure 1

Figure 1
<p>Complete sensor architecture.</p>
Full article ">Figure 2
<p>SNR distributions before and after noise reduction.</p>
Full article ">Figure 3
<p>The proposed 1D CNN architecture.</p>
Full article ">Figure 4
<p>Comparative SNR improvement across noise reduction techniques for SiNW-FET biosensor signals.</p>
Full article ">Figure 5
<p>FPGA-based 1D CNN accelerator design.</p>
Full article ">
13 pages, 1097 KiB  
Article
Research on the Application of Silver Nanowire-Based Non-Magnetic Transparent Heating Films in SERF Magnetometers
by Yi Ge, Yuhan Li, Yang Li, Xuejing Liu, Xiangmei Dong and Xiumin Gao
Sensors 2025, 25(1), 234; https://doi.org/10.3390/s25010234 - 3 Jan 2025
Viewed by 323
Abstract
We propose a non-magnetic transparent heating film based on silver nanowires (Ag-NWs) for application in spin-exchange relaxation-free (SERF) magnetic field measurement devices. To achieve ultra-high sensitivity in atomic magnetometers, the atoms within the alkali metal vapor cell must be maintained in a stable [...] Read more.
We propose a non-magnetic transparent heating film based on silver nanowires (Ag-NWs) for application in spin-exchange relaxation-free (SERF) magnetic field measurement devices. To achieve ultra-high sensitivity in atomic magnetometers, the atoms within the alkali metal vapor cell must be maintained in a stable and uniform high-temperature environment. Ag-NWs, as a transparent conductive material with exceptional electrical conductivity, are well suited for this application. By employing high-frequency AC heating, we effectively minimize associated magnetic noise. The experimental results demonstrate that the proposed heating film, utilizing a surface heating method, can achieve temperatures exceeding 140 °C, which is sufficient to vaporize alkali metal atoms. The average magnetic flux coefficient of the heating film is 0.1143 nT/mA. Typically, as the current increases, a larger magnetic field is generated. When integrated with the heating system discussed in this paper, this characteristic can effectively mitigate low-frequency magnetic interference. In comparison with traditional flexible printed circuits (FPC), the Ag-NWs heating film exhibits a more uniform temperature distribution. This magnetically transparent heating film, leveraging Ag-NWs, enhances atomic magnetometry and presents opportunities for use in chip-level gyroscopes, atomic clocks, and various other atomic devices. Full article
(This article belongs to the Section Optical Sensors)
Show Figures

Figure 1

Figure 1
<p>Non-magnetic electric heating structure.</p>
Full article ">Figure 2
<p>Ag-NWs’ transparent heating film structure.</p>
Full article ">Figure 3
<p>Simulation results of the interior walls and surrounding components of the alkali metal chamber. (<b>a</b>) Heating the alkali metal chamber using FPC heating film; (<b>b</b>) heating the alkali metal chamber using Ag-NWs’ transparent heating film; (<b>c</b>) FPC heating film structure; (<b>d</b>) Ag-NWs’ heating film structure.</p>
Full article ">Figure 4
<p>(<b>a</b>) Heating temperature of Ag-NWs’ transparent heating film at different voltages. (<b>b</b>) Simulation temperature curve (black) vs. experimental temperature curve (red).</p>
Full article ">Figure 5
<p>Alkali metal gas chamber temperature stability curve.</p>
Full article ">Figure 6
<p>(<b>a</b>) Simulation results for 1, 2, 3, and 4 films of Ag-NWs’ transparent heating films. (<b>b</b>) Thermal maps of internal temperatures for 1, 2, 3, and 4 films of Ag-NWs’ transparent heating films.</p>
Full article ">Figure 7
<p>Curve of normalized transmitted intensity as a function of temperature.</p>
Full article ">
13 pages, 3172 KiB  
Communication
Flat-Band AC Transport in Nanowires
by Vicenta Sánchez and Chumin Wang
Nanomaterials 2025, 15(1), 38; https://doi.org/10.3390/nano15010038 - 29 Dec 2024
Viewed by 275
Abstract
The electronic states in flat bands possess zero group velocity and null charge mobility. Recently, flat electronic bands with fully localized states have been predicted in nanowires, when their hopping integrals between first, second, and third neighbors satisfy determined relationships. Experimentally, these relationships [...] Read more.
The electronic states in flat bands possess zero group velocity and null charge mobility. Recently, flat electronic bands with fully localized states have been predicted in nanowires, when their hopping integrals between first, second, and third neighbors satisfy determined relationships. Experimentally, these relationships can only be closely achieved under external pressures. In this article, we study the alternating current (AC) in such nanowires having nearly flat electronic bands by means of a new independent channel method developed for the Kubo–Greenwood formula including hopping integrals up to third neighbors. The results reveal a large AC conductivity sensitive to the boundary conditions of measurement, where the charge carriers resonate with the external electric field by oscillating around their localized positions. Full article
(This article belongs to the Special Issue Theoretical Calculation Study of Nanomaterials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>By means of a unitary transformation, a cubically structured nanowire with cross section of 3 × 4 atoms and hopping interactions up to third neighbors through <math display="inline"><semantics> <mrow> <mi>t</mi> <mo>,</mo> <mo> </mo> <msup> <mi>t</mi> <mo>′</mo> </msup> </mrow> </semantics></math>, and <math display="inline"><semantics> <msup> <mi>t</mi> <mo>″</mo> </msup> </semantics></math> is represented by 12 independent channels indexed by <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mi>α</mi> <mo>,</mo> <mi>β</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>, where the system and its leads are connected by the hopping integrals <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>c</mi> </msub> <mo>,</mo> <mo> </mo> <msub> <msup> <mi>t</mi> <mo>′</mo> </msup> <mi>c</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <msup> <mi>t</mi> <mo>″</mo> </msup> <mi>c</mi> </msub> </mrow> </semantics></math> originated from their structural interfaces.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>m</b>) Band width of independent channels <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <mi>α</mi> <mo>,</mo> <mi>β</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math>, numbered in each figure, versus the hopping integral ratio <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>/</mo> <mi>t</mi> </mrow> </mrow> </semantics></math> for a nanowire with cross section of <math display="inline"><semantics> <mrow> <mn>3</mn> <mo>×</mo> <mn>4</mn> </mrow> </semantics></math> atoms, being <math display="inline"><semantics> <mrow> <mo>Φ</mo> <mo> </mo> <mo>=</mo> <mrow> <mrow> <mo stretchy="false">(</mo> <msqrt> <mn>5</mn> </msqrt> <mo> </mo> <mo>+</mo> <mn>1</mn> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Electronic density of states (DOS) versus energy with <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> <mo> </mo> <mo>|</mo> <mo> </mo> <mi>t</mi> <mo> </mo> <mo>|</mo> </mrow> </semantics></math> for the same nanowire of <a href="#nanomaterials-15-00038-f002" class="html-fig">Figure 2</a> with <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>/</mo> <mi>t</mi> </mrow> <mo>=</mo> <mrow> <mn>1</mn> <mo>/</mo> <mrow> <msqrt> <mn>2</mn> </msqrt> </mrow> </mrow> </mrow> </semantics></math> (blue line) and <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.707</mn> </mrow> </semantics></math> (red line). (<b>b</b>) Magnification of DOS around <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <msqrt> <mn>2</mn> </msqrt> <mo>|</mo> <mo> </mo> <mi>t</mi> <mo> </mo> <mo>|</mo> </mrow> </semantics></math> including those of channels (3,1) (magenta line), (3,2) (orange line), (3,3) (green line) and (3,4) (cyan line) for <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mn>0.707</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Zero-temperature electrical conductivity <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>σ</mi> <mo stretchy="false">(</mo> <mi>μ</mi> <mo>,</mo> <mi mathvariant="sans-serif">ω</mi> <mo>,</mo> <mn>0</mn> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <mrow> <msub> <mi>σ</mi> <mi>P</mi> </msub> </mrow> </mrow> </mrow> </semantics></math> versus the chemical potential (<math display="inline"><semantics> <mi>μ</mi> </semantics></math>) and the frequency (<math display="inline"><semantics> <mi mathvariant="sans-serif">ω</mi> </semantics></math>) for the same nanowire of <a href="#nanomaterials-15-00038-f003" class="html-fig">Figure 3</a> with a hopping integral ratio <math display="inline"><semantics> <mrow> <mi>τ</mi> <mo> </mo> <mo>=</mo> <mo> </mo> <mrow> <msup> <mi>t</mi> <mo>′</mo> </msup> <mo>/</mo> <mi>t</mi> </mrow> <mo>=</mo> <mn>0.707</mn> </mrow> </semantics></math> and (<b>b</b>) its magnification around <math display="inline"><semantics> <mrow> <mi>μ</mi> <mo>=</mo> <msqrt> <mn>2</mn> </msqrt> <mo>|</mo> <mo> </mo> <mi>t</mi> <mo> </mo> <mo>|</mo> </mrow> </semantics></math> plotted in the logarithmic scale of frequency.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Zero-temperature electrical conductivity <math display="inline"><semantics> <mrow> <mi>σ</mi> <mo stretchy="false">(</mo> <mi>μ</mi> <mo>,</mo> <mi mathvariant="sans-serif">ω</mi> <mo>,</mo> <mn>0</mn> <mo stretchy="false">)</mo> </mrow> </semantics></math> as a function of the chemical potential (<span class="html-italic">μ</span>) and frequency (ω) for the nanowire illustrated in <a href="#nanomaterials-15-00038-f001" class="html-fig">Figure 1</a> with (<b>a</b>) <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>c</mi> </msub> <mo> </mo> <mo>=</mo> <mn>0.999999</mn> <mo> </mo> <mi>t</mi> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>c</mi> </msub> <mo> </mo> <mo>=</mo> <mn>0.7</mn> <mo> </mo> <mi>t</mi> </mrow> </semantics></math>. (<b>c</b>) Magnification of the density of states (DOS) as a function of energy (<span class="html-italic">E</span>) with <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>=</mo> <msup> <mrow> <mn>10</mn> </mrow> <mrow> <mo>−</mo> <mn>23</mn> </mrow> </msup> <mo> </mo> <mo>|</mo> <mo> </mo> <mi>t</mi> <mo> </mo> <mo>|</mo> </mrow> </semantics></math> around <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>=</mo> <mn>1.4144144277628045</mn> <mo>|</mo> <mo> </mo> <mi>t</mi> <mo> </mo> <mo>|</mo> </mrow> </semantics></math> for the case <math display="inline"><semantics> <mrow> <msub> <mi>t</mi> <mi>c</mi> </msub> <mo> </mo> <mo>=</mo> <mn>0.999999</mn> <mo> </mo> <mi>t</mi> </mrow> </semantics></math>.</p>
Full article ">
13 pages, 3352 KiB  
Article
The Addition of MoO3 or SiO2 Nano-/Microfillers Thermally Stabilized and Mechanically Reinforce the PVDF-HFP/PVP Polymer Composite Thin Films
by Urška Gradišar Centa, Anja Pogačnik Krajnc, Lidija Slemenik Perše, Matic Šobak and Mohor Mihelčič
Coatings 2024, 14(12), 1603; https://doi.org/10.3390/coatings14121603 - 21 Dec 2024
Viewed by 426
Abstract
The properties of thin polymer films are influenced by the size of the fillers, their morphology, the surface properties and their distribution/interaction in the polymer matrix. In this work, thin polymer composite films with MoO3 or SiO2 nano and micro fillers [...] Read more.
The properties of thin polymer films are influenced by the size of the fillers, their morphology, the surface properties and their distribution/interaction in the polymer matrix. In this work, thin polymer composite films with MoO3 or SiO2 nano and micro fillers in PVDF-HFP/PVP polymer matrix were successfully fabricated using the solvent casting method. The effects of different types, sizes and morphologies of the inorganic fillers on the crystallization of the PVDF-HFP polymer were investigated, as well as the effects on the thermal and mechanical properties of the composites. Scanning electron microscopy, ATR-FTIR spectroscopy, differential scanning calorimetry, nanoindentation and uniaxial mechanical tests were used for characterization. The results showed that MoO3 nanowires thermally stabilized the polymer matrix, induced crystallization of the PVDF-HFP polymer in all three polymorphs (α-, β-, γ-phase) and formed a geometrical network in the polymer matrix, resulting in the highest elastic moduli, hardness and Young’s modulus. Full article
Show Figures

Figure 1

Figure 1
<p>Samples of (<b>A</b>) polymer matrix and polymer composites: with (<b>B</b>) MoO<sub>3</sub> nanowires, (<b>C</b>) MoO<sub>3</sub> microparticles, (<b>D</b>) SiO<sub>2</sub> nanoparticles, (<b>E</b>) SiO<sub>2</sub> microparticles for characterization.</p>
Full article ">Figure 2
<p>SEM images of nano- and micro-particles used: (<b>a</b>) MoO<sub>3</sub> nanowires, (<b>b</b>) MoO<sub>3</sub> microparticles, (<b>c</b>) SiO<sub>2</sub> nanoparticles, (<b>d</b>) SiO<sub>2</sub> microparticles.</p>
Full article ">Figure 3
<p>SEM images of the surface of (<b>a</b>) PVDF-HFP/PVP polymer matrix and polymer composite thin films and with the addition of: (<b>b</b>) MoO<sub>3</sub> nanowires, (<b>c</b>) MoO<sub>3</sub> microparticles, (<b>d</b>) SiO<sub>2</sub> nanoparticles, (<b>e</b>) SiO<sub>2</sub> microparticles. The enlargement of the polymer domains is shown in the insert.</p>
Full article ">Figure 4
<p>ATR-FTIR curves for PVDF-HFP/PVP polymer matrix and polymer composites with MoO<sub>3</sub> nanowires, MoO<sub>3</sub> microparticles, SiO<sub>2</sub> nanoparticles, SiO<sub>2</sub> microparticles.</p>
Full article ">Figure 5
<p>DSC curves for polymer composites: (<b>a</b>) second heating step, (<b>b</b>) cooling step.</p>
Full article ">Figure 6
<p>Nanoindentation: (<b>a</b>) elastic modulus and (<b>b</b>) hardness of PVDF-HFP/PVP polymer blend and polymer composites with MoO<sub>3</sub> and SiO<sub>2</sub> nano- and micro-particles.</p>
Full article ">Figure 7
<p>Mechanical properties of the polymer nanocomposites samples obtained with uniaxial mechanical tests. (<b>a</b>) Stress-strain curves, (<b>b</b>) values of Young’s modulus for PVDF-HFP/PVP polymer blend and polymer composites with MoO<sub>3</sub> and SiO<sub>2</sub> nano- and micro-particles.</p>
Full article ">
14 pages, 4160 KiB  
Article
Selective CO2 Detection at Room Temperature with Polyaniline/SnO2 Nanowire Composites
by Gen Li, Muhammad Hilal, Hyojung Kim, Jiyeon Lee, Zhiyong Chen, Bin Li, Yunhao Cui, Jian Hou and Zhicheng Cai
Coatings 2024, 14(12), 1590; https://doi.org/10.3390/coatings14121590 - 19 Dec 2024
Viewed by 412
Abstract
In this study, tin oxide (SnO2)/polyaniline (PANI) composite nanowires (NWs) with varying amounts of PANI were synthesized for carbon dioxide (CO2) gas sensing at room temperature (RT, 25 °C). SnO2 NWs were fabricated via the vapor–liquid–solid (VLS) method, [...] Read more.
In this study, tin oxide (SnO2)/polyaniline (PANI) composite nanowires (NWs) with varying amounts of PANI were synthesized for carbon dioxide (CO2) gas sensing at room temperature (RT, 25 °C). SnO2 NWs were fabricated via the vapor–liquid–solid (VLS) method, followed by coating with PANI. CO2 sensing investigations revealed that the sensor with 186 μL PANI exhibited the highest response to CO2 at RT. Additionally, the optimized sensor demonstrated excellent selectivity for CO2, long-term stability, and reliable performance across different humidity levels. The enhanced sensing performance of the optimized sensor was attributed to the formation of SnO2-PANI heterojunctions and the optimal PANI concentration. This study underscores the potential of SnO2-PANI composites for CO2 detection at RT. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the preparation process for PANI/SnO<sub>2</sub> composites.</p>
Full article ">Figure 2
<p>SEM and TEM characterization of PANI/SnO<sub>2</sub> composites: (<b>a</b>–<b>d</b>) SEM images of SP0, SP1, SP2, and SP3, respectively; (<b>e</b>–<b>h</b>) magnified SEM images showing surface morphology of individual nanowires; (<b>i</b>) TEM image of a single SP2 nanowire; (<b>j</b>) magnified TEM image revealing lattice fringes of the SnO<sub>2</sub> core and PANI coating.</p>
Full article ">Figure 3
<p>XRD patterns of PANI/SnO<sub>2</sub> composites (SP0, SP1, SP2, and SP3) showing characteristic peaks of SnO<sub>2</sub> and PANI with varying polymer content.</p>
Full article ">Figure 4
<p>XPS analysis of PANI/SnO<sub>2</sub> composites (SP0, SP1, SP2, and SP3): (<b>a</b>) Sn 3d peaks, (<b>b</b>–<b>d</b>) C 1s spectrum of PANI, (<b>e</b>–<b>g</b>) N 1s spectrum of PANI, and (<b>h</b>–<b>k</b>) O 1s peaks highlighting oxygen states.</p>
Full article ">Figure 5
<p>Gas sensing performance of PANI/SnO<sub>2</sub> composites: (<b>a</b>) dynamic response transient curves under varying CO<sub>2</sub> concentrations, (<b>b</b>) response–recovery curves, (<b>c</b>) response as a function of CO<sub>2</sub> concentration, (<b>d</b>) response and recovery times, (<b>e</b>) selectivity for various gases, (<b>f</b>) repeatability over multiple cycles, (<b>g</b>) long-term stability, and (<b>h</b>) humidity effects on CO<sub>2</sub> response and baseline resistance.</p>
Full article ">Figure 6
<p>Schematic representation of the sensing mechanism in PANI/SnO<sub>2</sub> composites.</p>
Full article ">
21 pages, 8007 KiB  
Article
Machine Learning-Based Modeling of pH-Sensitive Silicon Nanowire (SiNW) for Ion Sensitive Field Effect Transistor (ISFET)
by Nabil Ayadi, Ahmet Lale, Bekkay Hajji, Jérôme Launay and Pierre Temple-Boyer
Sensors 2024, 24(24), 8091; https://doi.org/10.3390/s24248091 - 18 Dec 2024
Viewed by 529
Abstract
The development of ion-sensitive field-effect transistor (ISFET) sensors based on silicon nanowires (SiNW) has recently seen significant progress, due to their many advantages such as compact size, low cost, robustness and real-time portability. However, little work has been done to predict the performance [...] Read more.
The development of ion-sensitive field-effect transistor (ISFET) sensors based on silicon nanowires (SiNW) has recently seen significant progress, due to their many advantages such as compact size, low cost, robustness and real-time portability. However, little work has been done to predict the performance of SiNW-ISFET sensors. The present study focuses on predicting the performance of the silicon nanowire (SiNW)-based ISFET sensor using four machine learning techniques, namely multilayer perceptron (MLP), nonlinear regression (NLR), support vector regression (SVR) and extra tree regression (ETR). The proposed ML algorithms are trained and validated using experimental measurements of the SiNW-ISFET sensor. The results obtained show a better predictive ability of extra tree regression (ETR) compared to other techniques, with a low RMSE of 1 × 10−3 mA and an R2 value of 0.9999725. This prediction study corrects the problems associated with SiNW -ISFET sensors. Full article
(This article belongs to the Section Electronic Sensors)
Show Figures

Figure 1

Figure 1
<p>The (ML) techniques used for the Ids prediction of the silicon nanowires ISFET.</p>
Full article ">Figure 2
<p>The Multi-Layer Perception Neural Networks (MLP).</p>
Full article ">Figure 3
<p>Program flow chart of Multilayer Perception (MLP).</p>
Full article ">Figure 4
<p>Architecture of Support Vector Regression (SVR).</p>
Full article ">Figure 5
<p>Program flow chart of the Support Vector Regression (SVR).</p>
Full article ">Figure 6
<p>Program flow chart of the Nonlinear Regression (NLR).</p>
Full article ">Figure 7
<p>Program flow chart of the Extra Trees regression (ETR).</p>
Full article ">Figure 8
<p>Comparison of SiNW-ISFET characteristics with varying wire lengths, obtained through four machine learning techniques and validated by measurements (Vds = 1 V): (<b>a</b>) SVR, (<b>b</b>) MLP, (<b>c</b>) NLR and (<b>d</b>) ETR.</p>
Full article ">Figure 9
<p>Prediction accuracy of four machine learning techniques with different lengths of nanowires: (<b>a</b>) SVR, (<b>b</b>) NLR, (<b>c</b>) MLP and (<b>d</b>) ETR.</p>
Full article ">Figure 10
<p>Comparison of the SiNW-ISFET characteristics with different numbers of wires, obtained through four machine learning techniques and validated by measurements (the nanowire length is 2 μm and Vds = 1 V): (<b>a</b>) SVR, (<b>b</b>) MLP, (<b>c</b>) NLR and (<b>d</b>) ETR.</p>
Full article ">Figure 11
<p>Prediction accuracy of four machine learning techniques with different numbers of wires: (<b>a</b>) SVR, (<b>b</b>) NLR, (<b>c</b>) MLP and (<b>d</b>) ETR.</p>
Full article ">Figure 12
<p>Comparison of the SiNW-ISFET characteristics with different gate lengths (0.73 μm and 3.73 μm), obtained through four machine learning techniques and validated by measurements (the nanowire length is 10 μm and Vds = 1 V): (<b>a</b>) SVR, (<b>b</b>) MLP, (<b>c</b>) NLR and (<b>d</b>) ETR.</p>
Full article ">Figure 13
<p>Prediction accuracy of four machine learning techniques with different gate lengths: (<b>a</b>) SVR, (<b>b</b>) NLR, (<b>c</b>) MLP and (<b>d</b>) ETR.</p>
Full article ">Figure 14
<p>Comparison of the SiNW-ISFET characteristics with different pH, obtained through four machine learning techniques and validated by measurements (Vds = 1 V): (<b>a</b>) SVR, (<b>b</b>) MLP, (<b>c</b>) NLR and (<b>d</b>) ETR.</p>
Full article ">Figure 15
<p>Prediction accuracy of four machine learning techniques with different pH values: (<b>a</b>) SVR, (<b>b</b>) NLR, (<b>c</b>) MLP and (<b>d</b>) ETR.</p>
Full article ">Figure 16
<p>Comparison of the sensitivities of SiNW-ISFET sensor, obtained through four machine learning techniques and validated by measurements (Ids = 200 µA and Vds = 1 V): (<b>a</b>) SVR, (<b>b</b>) MLP, (<b>c</b>) NLR and (<b>d</b>) ETR.</p>
Full article ">
17 pages, 6438 KiB  
Article
Synthesis and Study of Oxide Semiconductor Nanoheterostructures in SiO2/Si Track Template
by Alma Dauletbekova, Diana Junisbekova, Zein Baimukhanov, Aivaras Kareiva, Anatoli I. Popov, Alexander Platonenko, Abdirash Akilbekov, Ainash Abdrakhmetova, Gulnara Aralbayeva, Zhanymgul Koishybayeva and Jonibek Khamdamov
Crystals 2024, 14(12), 1087; https://doi.org/10.3390/cryst14121087 - 18 Dec 2024
Viewed by 639
Abstract
In this study, chemical deposition was used to synthesize structures of Ga2O3 -NW/SiO2/Si (NW—nanowire) at 348 K and SnO2-NW/SiO2/Si at 323 K in track templates SiO2/Si (either n- or p-type). The resulting [...] Read more.
In this study, chemical deposition was used to synthesize structures of Ga2O3 -NW/SiO2/Si (NW—nanowire) at 348 K and SnO2-NW/SiO2/Si at 323 K in track templates SiO2/Si (either n- or p-type). The resulting crystalline nanowires were δ-Ga2O3 and orthorhombic SnO2. Computer modeling of the delta phase of gallium oxide yielded a lattice parameter of a = 9.287 Å, which closely matched the experimental range of 9.83–10.03 Å. The bandgap is indirect with an Eg = 5.5 eV. The photoluminescence spectra of both nanostructures exhibited a complex band when excited by light with λ = 5.16 eV, dominated by luminescence from vacancy-type defects. The current–voltage characteristics of δ-Ga2O3 NW/SiO2/Si-p showed one-way conductivity. This structure could be advantageous in devices where a reverse current is undesirable. The p-n junction with a complex structure was formed. This junction consists of a polycrystalline nanowire base exhibiting n-type conductivity and a monocrystalline Si substrate with p-type conductivity. The I–V characteristics of SnO2-NW/SiO2/Si suggested near-metallic conductivity due to the presence of metallic tin. Full article
(This article belongs to the Section Inorganic Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>Electronic and nuclear energy losses for the 200 MeV Xe ion in the SiO<sub>2</sub>/Si structure calculated by the SRIM code.</p>
Full article ">Figure 2
<p>SEM image of surface of track template a-SiO<sub>2</sub>/Si.</p>
Full article ">Figure 3
<p>Current–voltage characteristic (CVC) measurement circuit.</p>
Full article ">Figure 4
<p>SEM image of the surface of the p-type template after CD of Ga<sub>2</sub>O<sub>3</sub> (t = 15 min), at T = 348 K.</p>
Full article ">Figure 5
<p>X-ray diffractograms of SiO<sub>2</sub>/Si-p templates after chemical deposition (CD) are as follows: (N1) with a deposition time of t = 15 min at T = 348 K (75 °C); (N2) with a deposition time of t = 30 min at T = 348 K.</p>
Full article ">Figure 6
<p>X-ray diffractograms of SiO<sub>2</sub>/Si-n type templates after chemical deposition (CD) at T = 348 K (75 °C) were obtained for two different deposition durations: (N3) for a deposition time of 15 min; (N4) for a deposition time of 30 min.</p>
Full article ">Figure 7
<p>Crystallographic cell of δ-Ga<sub>2</sub>O<sub>3</sub> consisting of 80 atoms. Gallium atoms are shown in brown, and oxygen is shown in red.</p>
Full article ">Figure 8
<p>Band structure and density of states of δ-Ga<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 9
<p>The SEM image of the SiO<sub>2</sub>/Si-n template surface after SnO<sub>2</sub> CD for 20 min at a temperature of 323 K.</p>
Full article ">Figure 10
<p>X-ray diffractogram of SnO<sub>2</sub> samples obtained by chemical deposition for 20 min at a temperature of 323 K.</p>
Full article ">Figure 11
<p>The PL spectrum of δ-Ga<sub>2</sub>O<sub>3</sub>-NW/SiO<sub>2</sub>/Si obtained by (CD) at T = 348 K (75 °C) with a deposition time of 15 min. The excitation was carried out using light with a wavelength of λ = 5.16 eV. The hollow circle line represents the experimental curve, while the red line indicates the set of Gauss components.</p>
Full article ">Figure 12
<p>The PL spectrum of SnO<sub>2</sub>-NW/SiO<sub>2</sub>/Si, excited at a wavelength of λ = 240 nm. The hollow circle line represents the experimental curve, while the blue line indicates the set of Gauss components.</p>
Full article ">Figure 13
<p>Current–voltage (I–V) characteristics of δ-Ga<sub>2</sub>O<sub>3</sub>-NW/SiO<sub>2</sub>/Si-p. The dashed curve represents the original, untreated sample. The red curve corresponds to the sample with δ-Ga<sub>2</sub>O<sub>3</sub> deposited for 15 min, a p-type Si substrate. The blue curve corresponds to the sample with δ-Ga<sub>2</sub>O<sub>3</sub> deposited for 30 min, a p-type Si substrate.</p>
Full article ">Figure 14
<p>Current–voltage (I–V) characteristics of SnO<sub>2</sub>-NW/SiO<sub>2</sub>/Si-n. The solid curve represents the original, untreated sample. The dashed curve represents the sample after CD with a deposition time of 20 min.</p>
Full article ">
18 pages, 5625 KiB  
Article
Study of Graphene Oxide and Silver Nanowires Interactions and Its Association with Electromagnetic Shielding Effectiveness
by Mila Milenković, Warda Saeed, Muhammad Yasir, Dusan Sredojević, Milica Budimir, Andjela Stefanović, Danica Bajuk-Bogdanović and Svetlana Jovanović
Int. J. Mol. Sci. 2024, 25(24), 13401; https://doi.org/10.3390/ijms252413401 - 13 Dec 2024
Viewed by 535
Abstract
Technological development has led to the need for materials able to block electromagnetic waves (EMWs) emitted from various devices. EMWs could negatively affect the working performance and lifetime of multiple instruments and measuring devices. New EMW shielding materials are being developed, while among [...] Read more.
Technological development has led to the need for materials able to block electromagnetic waves (EMWs) emitted from various devices. EMWs could negatively affect the working performance and lifetime of multiple instruments and measuring devices. New EMW shielding materials are being developed, while among nanomaterials, graphene-based composites have shown promising features. Herein, we have produced graphene oxide (GO), silver nanowires (AgNWs) composites, by varying the mass ratios of each component. UV-Vis, infrared, Raman spectroscopies, and thermogravimetric analysis proved the establishment of the interactions between them. For the first time, the strength and the nature of the interaction between GO sheets with various levels of oxidation and AgNWs were investigated using density function theory (DFT). The interaction energy between ideal graphene and AgNWs was calculated to be −48.9 kcal/mol, while for AgNWs and GO, this energy is almost doubled at −81.9 kcal/mol. The DFT results confirmed the interfacial polarization at the heterointerface via charge transfer and accumulation at the interface, improving the efficacy of EMW shielding. Our results indicated that AgNWs create a compact complex with GO due to charge transfer between them. Charge redistributions in GO-AgNWs composites resulted in an improved ability of the composite to block EMWs compared to GO alone. Full article
(This article belongs to the Section Physical Chemistry and Chemical Physics)
Show Figures

Figure 1

Figure 1
<p>UV-Vis spectra of GO and AgNWs (<b>a</b>), GO-AgNWs 5:5, GO-AgNWs 3:7, and GO-AgNWs 1:9 (<b>b</b>).</p>
Full article ">Figure 2
<p>Thermograms of GO-AgNWs 5:5 and rGO-AgNWs 5:5 (<b>a</b>), GO-AgNWs 3:7 and rGO-AgNWs 3:7 (<b>b</b>), and GO-AgNWs 1:9 and rGO-AgNWs 1:9 (<b>c</b>).</p>
Full article ">Figure 3
<p>FTIR spectra of GO, rGO, GO-AgNWs 5:5, rGO-AgNWs 5:5, GO-AgNWs 3:7, rGO-AgNWs 3:7, GO-AgNWs 1:9, and rGO-AgNWs 1:9.</p>
Full article ">Figure 4
<p>Raman spectra of GO, GO-AgNWs 5:5, GO-AgNWs 3:7, GO-AgNWs 1:9, (<b>a</b>) rGO, rGO-AgNWs 5:5, rGO-AgNWs 3:7, and rGO-AgNWs 1:9 (<b>b</b>).</p>
Full article ">Figure 5
<p>The optimized structure (<b>a</b>), natural bond orbital (NBO) charges (<b>b</b>), the molecular electrostatic potentials (MEP) (<b>c</b>), and the total density of states of the Ag30 cluster (<b>d</b>). Green and maroon denote positive and negative regions of the wavefunction.</p>
Full article ">Figure 6
<p>The optimized structures of Ag<sub>30</sub>@C<sub>40</sub>H<sub>16</sub> (<b>a</b>), Ag<sub>30</sub>@C<sub>40</sub>H<sub>16</sub>O<sub>4</sub> (<b>b</b>), Ag<sub>30</sub>@C<sub>40</sub>H<sub>16</sub>(OH)<sub>4</sub> (<b>c</b>), and Ag<sub>30</sub>@C<sub>40</sub>H<sub>16</sub>O<sub>2</sub>(OH)<sub>2</sub> (<b>d</b>) adducts as calculated at the B3LYP-D3/6-31G(d,p)/LANL2DZ level.</p>
Full article ">Figure 7
<p>MEP plots of Ag<sub>30</sub>@C<sub>40</sub>H<sub>16</sub> cluster (<b>a</b>), [Ag<sub>30</sub>@C<sub>40</sub>H<sub>15</sub>O<sub>2</sub>-COO]<sup>−</sup> (<b>b</b>), [Ag<sub>30</sub>@C<sub>40</sub>H<sub>14</sub>(OH)<sub>2</sub>-COO-O]<sup>2−</sup> (<b>c</b>), and [Ag<sub>30</sub>@C<sub>40</sub>H<sub>15</sub>O<sub>2</sub>(OH)<sub>2</sub>-COO]<sup>−</sup> (<b>d</b>) adducts.</p>
Full article ">Figure 8
<p>Top and side views of free-standing films recorded using SEM of GO (<b>a</b>), GO-AgNWs 5:5 (<b>b</b>), GO-AgNWs 3:7 (<b>c</b>), and GO-AgNWs 1:9 (<b>d</b>).</p>
Full article ">Figure 9
<p>SE<sub>T</sub>, SE<sub>A</sub>, and SE<sub>R</sub> values for GO-AgNWs 5:5 (<b>a</b>), GO-AgNWs 3:7 (<b>b</b>), and GO-AgNWs 1:9 (<b>c</b>), measured in the frequency range of 8–12 GHz.</p>
Full article ">
9 pages, 2796 KiB  
Article
Luminescent Nanocrystal Probes for Monitoring Temperature and Thermal Energy Dissipation of Electrical Microcircuit
by Dawid Jankowski, Kamil Wiwatowski, Michał Żebrowski, Aleksandra Pilch-Wróbel, Artur Bednarkiewicz, Sebastian Maćkowski and Dawid Piątkowski
Nanomaterials 2024, 14(24), 1985; https://doi.org/10.3390/nano14241985 - 11 Dec 2024
Viewed by 564
Abstract
In this work, we present an experimental approach for monitoring the temperature of submicrometric, real-time operating electrical circuits using luminescence thermometry. For this purpose, we utilized lanthanide-doped up-converting nanocrystals as nanoscale temperature probes, which, combined with a highly sensitive confocal photoluminescence microscope, enabled [...] Read more.
In this work, we present an experimental approach for monitoring the temperature of submicrometric, real-time operating electrical circuits using luminescence thermometry. For this purpose, we utilized lanthanide-doped up-converting nanocrystals as nanoscale temperature probes, which, combined with a highly sensitive confocal photoluminescence microscope, enabled temperature monitoring with spatial resolution limited only by the diffraction of light. To validate our concept, we constructed a simple model of an electrical microcircuit based on a single silver nanowire with a diameter of approximately 100 nm and a length of about 50 µm, whose temperature increase was induced by electric current flow. By driving electric current only along one half of the nanowire, we created a dual-function microstructure, where one section is a resistive heater, while the other operates as a radiator. Such a combination realistically reflects the electronic circuit and its thermal behavior. We demonstrated that nanocrystals distributed around this circuit allow for remote temperature readout and enable precise monitoring of the thermal energy propagation and heat dissipation processes, which are crucial for designing and developing highly integrated electronic on-chip devices. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Illustration of the energy diagram and up-conversion mechanism in nanocrystals activated by Er<sup>3+</sup> and Yb<sup>3+</sup> ions. Note the small energy gap between the <sup>4</sup>S<sub>3/2</sub> and <sup>2</sup>H<sub>11/2</sub> levels, whose relative populations can be thermally modified. Excited populations of these levels are denoted as N* and N**, respectively. (<b>b</b>) Example emission spectrum of the examined nanocrystal dense layer at temperatures of 25 °C and 55 °C. The spectra were normalized at 545 nm (N).</p>
Full article ">Figure 2
<p>Conceptual diagram of the model electrical microcircuit illustrating the configuration of the heating element and the radiator. Objects are not to scale.</p>
Full article ">Figure 3
<p>A simplified block diagram of the experimental setup. The components are labeled as follows: EC—electrical contact; DM—dichroic mirror; PH—diaphragm (pinhole); SF—short-pass filter; FM—flip mirror; BF—band-pass filter; and SPCM—single-photon counting module.</p>
Full article ">Figure 4
<p>(<b>a</b>) A selected silver nanowire imaged in backscattered laser light, with the heating element and radiator indicated, where <span class="html-fig-inline" id="nanomaterials-14-01985-i001"><img alt="Nanomaterials 14 01985 i001" src="/nanomaterials/nanomaterials-14-01985/article_deploy/html/images/nanomaterials-14-01985-i001.png"/></span> represents electrical contacts. (<b>b</b>) The same sample area captured in photoluminescence mode, showing the distribution of luminescent probes (nanocrystals) and the temperature readout positions <span class="html-fig-inline" id="nanomaterials-14-01985-i002"><img alt="Nanomaterials 14 01985 i002" src="/nanomaterials/nanomaterials-14-01985/article_deploy/html/images/nanomaterials-14-01985-i002.png"/></span>. (<b>c</b>) A microscope image of the nanowire connected to an external power supply, where the laser spot (pink) indicates the position of the sample over the objective.</p>
Full article ">Figure 5
<p>(<b>a</b>) Luminescence spectra recorded at Point 1 (Point 4 in the inset) for different current intensities flowing through the heating element. The spectra were normalized at 545 nm (N). (<b>b</b>) Temperature increase recorded at all measurement points as a function of current intensity.</p>
Full article ">
10 pages, 7445 KiB  
Article
Controlling Poly(3-hexythiophene) Hierarchical Polymer/SWCNT Nanohybrid Shish-Kebab Morphologies in Marginal Solvents
by Kevin Schnittker, Zahra Bahrami and Joseph Andrews
Crystals 2024, 14(12), 1065; https://doi.org/10.3390/cryst14121065 - 10 Dec 2024
Viewed by 499
Abstract
In organic optoelectronic devices, the self-assembly behavior of the conjugated polymer poly(3-hexylthiophene) (P3HT) into structured aggregates significantly influences the device’s performance, with processing conditions playing a key role. Incorporating carbon nanotubes (CNTs) into a P3HT solution can form hierarchical supramolecular structures known as [...] Read more.
In organic optoelectronic devices, the self-assembly behavior of the conjugated polymer poly(3-hexylthiophene) (P3HT) into structured aggregates significantly influences the device’s performance, with processing conditions playing a key role. Incorporating carbon nanotubes (CNTs) into a P3HT solution can form hierarchical supramolecular structures known as nanohybrid shish-kebabs (NHSKs). These structures alter the morphology of polymer aggregates and provide an alternative pathway for improved charge transport in thin film devices. Herein, we investigated the impact of solvent quality using different combinations of chloroform and anisole during the quasi-isothermal crystallization of P3HT:CNTs. We found that NHSKs of different nanowire lengths can be formed through changing solvent quality while maintaining a constant P3HT:SWCNT ratio and a constant SWCNT concentration. Optical absorbance measurements showed that increasing the amount of the good solvent (chloroform) to 10.19% (v/v) reduced the exciton bandwidth by 36.4% compared to the NHSK solution that only contained ~2.37% (v/v). This observation demonstrates the importance of solvent quality and how this processing parameter directly leads to the enhanced crystallization of supramolecular structures. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of the NHSK solution crystallization process. (<b>A</b>) Dispersion of SWCNTs in chloroform and (<b>B</b>) NHSK solution cooling to room temperature. (Figure created with BioRender.com).</p>
Full article ">Figure 2
<p>Absorbance spectra of NHSK and neat P3HT solutions.</p>
Full article ">Figure 3
<p>SEM images of drop-casted NHSK solutions: (<b>A</b>) 0.05 mg/mL, (<b>B</b>) 0.1 mg/mL, (<b>C</b>) 0.25 mg/mL, and (<b>D</b>) 0.4 mg/mL.</p>
Full article ">Figure 4
<p>AFM images of drop-casted NHSK solutions: (<b>A</b>) 0.4 mg/mL, (<b>B</b>) 0.25 mg/mL, (<b>C</b>) 0.1 mg/mL, and (<b>D</b>) 0.05 mg/mL.</p>
Full article ">Figure 5
<p>TEM images of (<b>A</b>) 0.25 mg/mL NHSK sample, (<b>B</b>) 0.1 mg/mL NHSK sample, (<b>C</b>) neat P3HT nanofibers formed in anisole, and (<b>D</b>) neat SWCNTs.</p>
Full article ">Figure 6
<p>Nanowire length distribution between the 0.25 mg/mL and 0.1 mg/mL NHSK samples. Width measurements of the neat P3HT nanofibers are included for comparison. Measurements are taken from TEM images.</p>
Full article ">
16 pages, 3015 KiB  
Article
A Low-Cost Flexible Optoelectronic Synapse Based on ZnO Nanowires for Neuromorphic Computing
by Yongqing Yue, Zixia Yu, Fangpei Li, Wenbo Peng, Quanzhe Zhu and Yongning He
Sensors 2024, 24(23), 7788; https://doi.org/10.3390/s24237788 - 5 Dec 2024
Viewed by 630
Abstract
Neuromorphic computing, inspired by the brain, holds significant promise for advancing artificial intelligence. Artificial optoelectronic synapses, which can convert optical signals into electrical signals, play a crucial role in neuromorphic computing. In this study, we successfully fabricated a flexible artificial optoelectronic synapse device [...] Read more.
Neuromorphic computing, inspired by the brain, holds significant promise for advancing artificial intelligence. Artificial optoelectronic synapses, which can convert optical signals into electrical signals, play a crucial role in neuromorphic computing. In this study, we successfully fabricated a flexible artificial optoelectronic synapse device based on the ZnO/PDMS structure by utilizing the magnetron sputtering technique to deposit the ZnO film on a flexible substrate. Under UV light illumination, the device exhibits excellent synaptic plasticity, including excitatory postsynaptic current (EPSC), short-term potentiation (STP), and paired-pulse facilitation (PPF). By growing ZnO nanowires, we improved the fabrication processes and further enhanced the synaptic properties of the device, demonstrating long-term potentiation (LTP) and the transition from short-term memory (STM) to long-term memory (LTM). Additionally, the device exhibits outstanding flexibility, maintaining stable synaptic plasticity under bending conditions. This device shows broad application potential in mimicking visual systems and is expected to contribute significantly to the development of neuromorphic computing. Full article
Show Figures

Figure 1

Figure 1
<p>PDMS flexible substrate ZnO artificial optoelectronic synapse device. (<b>a</b>) Main fabrication processes. (<b>b</b>) Device structure and testing platform. (<b>c</b>) SEM images of the ZnO thin film, including the side view and top view (upper right). (<b>d</b>) Transmission spectrum. (<b>e</b>) PL spectrum.</p>
Full article ">Figure 2
<p>(<b>a</b>) I–V characteristics under varying UV light intensities. (<b>b</b>) Transient response curves under varying UV light intensities. (<b>c</b>) Responsivity (<span class="html-italic">R</span>) and (<b>d</b>) detectivity (<span class="html-italic">D</span>*) with a bias of +5 V under varying UV light power densities.</p>
Full article ">Figure 3
<p>(<b>a</b>) Switching characteristics under varying UV light intensities. (<b>b</b>) Transient response curves for different illumination durations at the same UV light intensity. (<b>c</b>) The PPF variation with varying illumination Δt. (<b>d</b>) I–t curves corresponding to each Δt.</p>
Full article ">Figure 4
<p>Transient response curves of devices under varying UV light intensities, and SEM images of side view and top view for different nanowire growth conditions. Growth solution concentration of (<b>a</b>) 25 mmol/L, ammonia-free; (<b>b</b>) 50 mmol/L, no ammonia; (<b>c</b>) 100 mmol/L, no ammonia; (<b>d</b>) 25 mmol/L, ammonia-containing.</p>
Full article ">Figure 5
<p>The PPF variation with varying illumination Δt and I–t curves corresponding to each Δt (upper right) for devices under different nanowire growth conditions. Growth solution concentration of (<b>a</b>) 25 mmol/L, ammonia-free; (<b>b</b>) 50 mmol/L, no ammonia; (<b>c</b>) 100 mmol/L, no ammonia; (<b>d</b>) 25 mmol/L, ammonia-containing.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic diagram of the bending experiment. Transient response curves under varying UV light intensities for the device with the nanowire growth solution concentration of 100 mmol/L in various bending states. (<b>b</b>) Initial state. (<b>c</b>) Cylinder rod diameter d = 4 cm. (<b>d</b>) Cylinder rod diameter d = 3 cm. (<b>e</b>) Cylinder rod diameter d = 2 cm. (<b>f</b>) Recovery to the initial state after bending.</p>
Full article ">Figure 7
<p>(<b>a</b>) The 3 × 3 flexible optoelectronic synapse array and the patterned mask. The UV light irradiates the mask and then illuminate the unshielded devices in the array. (<b>b</b>) Images of the array after 30 s of UV light illumination, and following 30 s and 120 s of forgetting. (The color intensity reflects the logarithmic change in device current).</p>
Full article ">
Back to TopTop