[go: up one dir, main page]

Next Issue
Volume 12, August-1
Previous Issue
Volume 12, July-1
 
 

Nanomaterials, Volume 12, Issue 14 (July-2 2022) – 195 articles

Cover Story (view full-size image): Fluorescent defects hosted by 2D hexagonal boron nitride (hBN) feature bright quantum emission at room temperature over a wide spectral range. Due to this broad spectral coverage, hBN quantum emitters are a promising candidate for generating quantum light for future quantum networks, where different hybrid quantum systems are interfaced with each other. Here, Cholsuk et al. report on how these hBN quantum emitters can be tailored for use in modern quantum technologies. With theoretical simulations, a large database of active emitters is established at low-loss wavelengths for quantum communication or with efficient coupling to state-of-the-art solid-state qubits and quantum memories. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
17 pages, 4624 KiB  
Article
Synthesis and Characterization of Hematite-Based Nanocomposites as Promising Catalysts for Indigo Carmine Oxidation
by Andrei Cristian Kuncser, Arpad Mihai Rostas, Rodica Zavoianu, Octavian Dumitru Pavel, Ioana Dorina Vlaicu, Mihaela Badea, Daniela Cristina Culita, Alina Tirsoaga and Rodica Olar
Nanomaterials 2022, 12(14), 2511; https://doi.org/10.3390/nano12142511 - 21 Jul 2022
Cited by 4 | Viewed by 2028
Abstract
The hematite-based nanomaterials are involved in several catalytic organic and inorganic processes, including water decontamination from organic pollutants. In order to develop such species, a series of bimetallic hematite-based nanocomposites were obtained by some goethite composites-controlled calcination. Their composition consists of various phases [...] Read more.
The hematite-based nanomaterials are involved in several catalytic organic and inorganic processes, including water decontamination from organic pollutants. In order to develop such species, a series of bimetallic hematite-based nanocomposites were obtained by some goethite composites-controlled calcination. Their composition consists of various phases such as α-FeOOH, α-Fe2O3 or γ-Fe2O3 combined with amorphous (Mn2O3, Co3O4, NiO, ZnO) or crystalline (CuO) oxides of the second transition ion from the structure. The component dimensions, either in the 10–30 or in the 100–200 nm range, together with the quasi-spherical or nanorod-like shapes, were provided by Mössbauer spectroscopy and powder X-ray diffraction as well as transmission electron microscopy data. The textural characterization showed a decrease in the specific area of the hematite-based nanocomposites compared with corresponding goethites, with the pore volume ranging between 0.219 and 0.278 cm3g−1. The best catalytic activity concerning indigo carmine removal from water in hydrogen peroxide presence was exhibited by a copper-containing hematite-based nanocomposite sample that reached a dye removal extent of over 99%, which correlates with both the base/acid site ratio and pore size. Moreover, Cu-hbnc preserves its catalytic activity even after four recyclings, when it still reached a dye removal extent higher than 90%. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structure of indigo carmine.</p>
Full article ">Figure 2
<p>Temperature dependence of the <b>Fe-hbnc</b> (<b>a</b>), <b>Mn-hbnc</b> (<b>b</b>), <b>Co-hbnc</b> (<b>c</b>), <b>Ni-hbnc</b> (<b>d</b>), <b>Cu-hbnc</b> (<b>e</b>), and <b>Zn-hbnc</b> (<b>f</b>) EPR signals.</p>
Full article ">Figure 3
<p>Temperature dependence of the EPR signal peak-to-peak linewidth (ΔLW<sub>PP</sub>) for all species.</p>
Full article ">Figure 4
<p>TEM results: CTEM, HRTEM, STEM, and EDS mappings for <b>Fe-hbnc</b> (<b>a</b>), <b>Mn-hbnc</b> (<b>b</b>), <b>Co-hbnc</b> (<b>c</b>), <b>Ni-hbnc</b> (<b>d</b>), Cu-<b>hbnc</b> (<b>e</b>), and <b>Zn-hbnc</b> (<b>f</b>).</p>
Full article ">Figure 5
<p>N<sub>2</sub> adsorption–desorption isotherms and pore size distribution (inset) of the samples: <b>Fe−hbnc</b> (<b>a</b>), <b>Mn−hbnc</b> (<b>b</b>), <b>Co−hbnc (c)</b>, <b>Ni−hbnc</b> (<b>d</b>), <b>Cu−hbnc</b> (<b>e</b>), and <b>Zn−hbnc</b> (<b>f</b>).</p>
Full article ">Figure 6
<p>Linear correlation of the DR % and the ratio base/acid sites in the catalyst samples ((<b>A</b>)—IC initial concentration 0.03 mM, 25 °C, 2 h, 150 rpm, 1 wt% catalyst; (<b>B</b>)—molar ratio H<sub>2</sub>O<sub>2</sub>/IC = 32.6, 25 °C, 2 h, 150 rpm, 1 wt% catalyst).</p>
Full article ">Figure 7
<p>DR % during five reaction cycles on <b>Fe-hbnc</b> and <b>Cu-hbnc</b> (IC initial concentration 0.03 mM, molar ratio H<sub>2</sub>O<sub>2</sub>/IC = 32.6, 25 °C, 2 h, 150 rpm, 1 wt% catalyst).</p>
Full article ">
14 pages, 6931 KiB  
Article
Fabrication of Durable Superhydrophobic Surface for Versatile Oil/Water Separation Based on HDTMS Modified PPy/ZnO
by Shumin Fan, Sujie Jiang, Zhenjie Wang, Pengchao Liang, Wenxiu Fan, Kelei Zhuo and Guangri Xu
Nanomaterials 2022, 12(14), 2510; https://doi.org/10.3390/nano12142510 - 21 Jul 2022
Cited by 6 | Viewed by 2254
Abstract
Superhydrophobic materials have been widely applied in rapid removal and collection of oils from oil/water mixtures for increasing damage to environment and human beings caused by oil-contaminated wastewater and oil spills. Herein, superhydrophobic materials were fabricated by a novel polypyrrole (PPy)/ZnO coating followed [...] Read more.
Superhydrophobic materials have been widely applied in rapid removal and collection of oils from oil/water mixtures for increasing damage to environment and human beings caused by oil-contaminated wastewater and oil spills. Herein, superhydrophobic materials were fabricated by a novel polypyrrole (PPy)/ZnO coating followed by hexadecyltrimethoxysilane (HDTMS) modification for versatile oil/water separation with high environmental and excellent reusability. The prepared superhydrophobic surfaces exhibited water contact angle (WCA) greater than 150° and SA less than 5°. The superhydrophobic fabric could be applied for separation of heavy oil or light oil/water mixtures and emulsions with the separation efficiencies above 98%. The coated fabric also realized highly efficient separation with harsh environmental solutions, such as acid, alkali, salt, and hot water. The superhydrophobic fabric still remained, even after 80 cycles of separation and 12 months of storage in air, proving excellent durability. These novel superhydrophobic materials have indicated great development potentials for oil/water separation in practical applications. Full article
(This article belongs to the Special Issue Advanced Nanomaterials for Environmental Remediation)
Show Figures

Figure 1

Figure 1
<p>SEM images of raw and PPy/ZnO/HDTMS coated samples. (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>,<b>p</b>): corn stalk, sawdust, cotton, fabric, sponge, copper mesh; (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>,<b>q</b>): the coated samples; (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>,<b>r</b>): higher magnification images of the coated samples.</p>
Full article ">Figure 2
<p>Photographs of original (<b>left</b>) and PPy/ZnO/HDTMS coated samples (<b>right</b>). (<b>a</b>,<b>b</b>): corn stalk; (<b>c</b>,<b>d</b>): sawdust; (<b>e</b>,<b>f</b>): cotton; (<b>g</b>,<b>h</b>): fabric; (<b>i</b>,<b>j</b>): sponge; (<b>k</b>,<b>l</b>): copper mesh.</p>
Full article ">Figure 3
<p>FTIR spectrum.</p>
Full article ">Figure 4
<p>XRD patterns.</p>
Full article ">Figure 5
<p>(<b>a</b>) Reaction steps of the PPy/ZnO/HDTMS coating; (<b>b</b>) A reaction mechanism diagram for HDTMS grafting on ZnO nanoparticles.</p>
Full article ">Figure 6
<p>Oil/water separation. (<b>a</b>), tetrachloromethane/water mixture; (<b>b</b>), petroleum ether/water mixture; (<b>c</b>), water-in-tetrachloromethane emulsion; (<b>d</b>), the emulsion under optical microscope before (<b>left</b>) and after (<b>right</b>) separation.</p>
Full article ">Figure 7
<p>The resistance to corrosive solutions of the coated fabric. (<b>a</b>) The WCA of 5 μL of hot water (100 °C), HCl solution (1M), NaOH solution (1M), and NaCl solution (1M) on coated fabric; (<b>b</b>) The oil filtrate fraction of tetrachloromethane/corrosive solution mixtures.</p>
Full article ">Figure 8
<p>Durability of the coated fabric. (<b>a</b>) The oil filtrate fraction with number of cycles; (<b>b</b>) The SEM images of the coated fabric after being reused for 80 cycles; (<b>c</b>) The SEM images of the coated fabric after 80 abrasion cycles; (<b>d</b>) The WCA of coated fabric with abrasion cycles; (<b>e</b>) The WCA of coated fabric with the storage time in air.</p>
Full article ">Figure 8 Cont.
<p>Durability of the coated fabric. (<b>a</b>) The oil filtrate fraction with number of cycles; (<b>b</b>) The SEM images of the coated fabric after being reused for 80 cycles; (<b>c</b>) The SEM images of the coated fabric after 80 abrasion cycles; (<b>d</b>) The WCA of coated fabric with abrasion cycles; (<b>e</b>) The WCA of coated fabric with the storage time in air.</p>
Full article ">
15 pages, 5898 KiB  
Article
Self-Supply Oxygen ROS Reactor via Fenton-like Reaction and Modulating Glutathione for Amplified Cancer Therapy Effect
by Huanli Zhang, Wei Ma, Zhiqiang Wang, Xiaodan Wu, Hui Zhang, Wen Fang, Rui Yan and Yingxue Jin
Nanomaterials 2022, 12(14), 2509; https://doi.org/10.3390/nano12142509 - 21 Jul 2022
Cited by 16 | Viewed by 2646
Abstract
Reactive oxygen species (ROS) are highly reactive oxidant molecules that can kill cancer cells through irreversible damage to biomacromolecules. ROS-mediated cancer therapies, such as chemodynamic (CDT) and photodynamic therapy (PDT), are often limited by the hypoxia tumor microenvironment (TME) with high glutathione (GSH) [...] Read more.
Reactive oxygen species (ROS) are highly reactive oxidant molecules that can kill cancer cells through irreversible damage to biomacromolecules. ROS-mediated cancer therapies, such as chemodynamic (CDT) and photodynamic therapy (PDT), are often limited by the hypoxia tumor microenvironment (TME) with high glutathione (GSH) level. This paper reported the preparation, characterization, in vitro and in vivo antitumor bioactivity of a meso-tetra(4-carboxyphenyl)porphine (TCPP)-based therapeutic nanoplatform (CMMFTP) to overcome the limitations of TME. Using Cu2+ as the central ion and TCPP as the ligand, the 2D metal-organic framework Cu-TCPP was synthesized by the solvothermal method, then CMMFTP was prepared by modifying MnO2, folic acid (FA), triphenylphosphine (TPP), and poly (allylamine hydrochloride) (PAH) on the surface of Cu-TCPP MOFs. CMMFTP was designed as a self-oxygenating ROS nanoreactor based on the PDT process of TCPP MOFs and the CDT process by Cu(II) and MnO2 components (mainly through Fenton-like reaction). The in vitro assay suggested CMMFTP caused a 96% lethality rate against Hela cells (MTT analysis) in specific response to TME stimulation. Moreover, the Cu(II) and MnO2 in CMMFTP efficiently depleted the glutathione (80%) in tumor cells and consequently amplified ROS levels to improve CDT/PDT effects. The FA-induced tumor targeting and TPP-induced mitochondria targeting further enhanced the antitumor activity. Therefore, the nanoreactor based on dual targeting and self-oxygenation-enhanced ROS mechanism provided a new strategy for cancer therapy. Full article
(This article belongs to the Section Biology and Medicines)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Transmission electron microscopy (TEM) image of CMMFTP; (<b>B</b>) high-resolution transmission electron microscope (HR-TEM) image of CMMFTP (Inset: HR-TEM image of MnO<sub>2</sub> nanocrystals); (<b>C</b>) X-ray diffraction (XRD) patterns of CMMFTP; (<b>D</b>) X-ray photoelectron spectroscopy (XPS) spectra of CMMFTP; (<b>E</b>) high-resolution X-ray photoelectron spectroscopy (HR-XPS) of C 1s; (<b>F</b>) HR-XPS of N 1s; (<b>G</b>) HR-XPS of Cu 2p; (<b>H</b>) HR-XPS of O 1s; (<b>I</b>) HR-XPS of Mn 2p; (<b>J</b>) Fourier transform infrared (FT-IR) spectra of TCPP, Cu-MOFs and CMMFTP; (<b>K</b>) UV-visible absorption spectra of TCPP, Cu-MOFs and CMMFTP; (<b>L</b>) Fluorescence emission spectrum of CMMFTP; (<b>M</b>) dark-field TEM image and corresponding elemental mappings images of CMMFTP.</p>
Full article ">Figure 2
<p>(<b>A</b>) Detection of <sup>•</sup>OH generation capacity of various samples using MB; (<b>B</b>) <sup>•</sup>OH detection by ESR signals under different conditions; (<b>C</b>) detection of O<sub>2</sub> production capacity of various samples using RDPP as a dissolved oxygen probe; (<b>D</b>) examine the ability of CMMFTP photosensitization to produce <sup>1</sup>O<sub>2</sub> using DPBF as singlet oxygen indicator probe; (<b>E</b>) standard curves of degradation rate and light time of DPBF; (<b>F</b>) fluorescence spectra and (<b>G</b>) fluorescence intensity histograms of <sup>1</sup>O<sub>2</sub> detected with the SOSG probe; (<b>H</b>) detection of GSH consumption by CMMFTP with DTNB probe.</p>
Full article ">Figure 3
<p>(<b>A</b>) Time (100 µg/mL, 0–90 min) dependence and (<b>B</b>) concentration (30 min, 0–400 µg/mL) dependence images of cell uptake; (<b>C</b>) detection of mitochondrial targeting of CMMFTP with Mito-Tracker Green; (<b>D</b>) detection of the intracellular GSH depletion capacity of CMMFTP; (<b>E</b>) dark toxicity of CMMFTP in Hela cells; (<b>F</b>) dark toxicity of CMMFTP in normal cells; (<b>G</b>–<b>I</b>) comparison of the cytotoxicity of Cu-MOFs, Cu-MOF@MnO<sub>2</sub>, CMMFTP under different conditions. <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>(<b>A</b>) Detection of intracellular ROS; (<b>B</b>) calcein-AM and PI co-staining of living and dead cells.</p>
Full article ">Figure 5
<p>(<b>A</b>) Photographs of tumor-bearing mice in different experimental groups after 14 days of treatment; (<b>B</b>) images of tumors after different treatments; (<b>C</b>) changes in body weight of mice during treatment; (<b>D</b>) tumor volume and (<b>E</b>) tumor weight in different experimental groups after 14 days of treatment; (<b>F</b>) H&amp;E staining images of heart, liver, spleen, lung, kidney, and tumors in different experimental groups after 14 days of treatment. Scale bar: 100 μm. <sup>∗</sup>, <span class="html-italic">p</span> &lt; 0.05; <sup>∗∗</sup>, <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Scheme 1
<p>The synthetic route and anti-tumor mechanism of a TME-responsive self-oxygenation-enhanced nanoreactor (CMMFTP) based on CDT/PDT synergy and GSH depletion strategy.</p>
Full article ">
9 pages, 2676 KiB  
Article
AlGaN Quantum Disk Nanorods with Efficient UV-B Emission Grown on Si(111) Using Molecular Beam Epitaxy
by Dongqi Zhang, Tao Tao, Haiding Sun, Siqi Li, Hongfeng Jia, Huabin Yu, Pengfei Shao, Zhenhua Li, Yaozheng Wu, Zili Xie, Ke Wang, Shibing Long, Bin Liu, Rong Zhang and Youdou Zheng
Nanomaterials 2022, 12(14), 2508; https://doi.org/10.3390/nano12142508 - 21 Jul 2022
Viewed by 1544
Abstract
AlGaN nanorods have attracted increasing amounts of attention for use in ultraviolet (UV) optoelectronic devices. Here, self-assembled AlGaN nanorods with embedding quantum disks (Qdisks) were grown on Si(111) using plasma-assisted molecular beam epitaxy (PA-MBE). The morphology and quantum construction of the nanorods were [...] Read more.
AlGaN nanorods have attracted increasing amounts of attention for use in ultraviolet (UV) optoelectronic devices. Here, self-assembled AlGaN nanorods with embedding quantum disks (Qdisks) were grown on Si(111) using plasma-assisted molecular beam epitaxy (PA-MBE). The morphology and quantum construction of the nanorods were investigated and well-oriented and nearly defect-free nanorods were shown to have a high density of about 2 × 1010 cm−2. By controlling the substrate temperature and Al/Ga ratio, the emission wavelengths of the nanorods could be adjusted from 276 nm to 330 nm. By optimizing the structures and growth parameters of the Qdisks, a high internal quantum efficiency (IQE) of the AlGaN Qdisk nanorods of up to 77% was obtained at 305 nm, which also exhibited a shift in the small emission wavelength peak with respect to the increasing temperatures during the PL measurements. Full article
(This article belongs to the Section Nanophotonics Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The schematic of the AlGaN Qdisk nanorods; (<b>b</b>) the schematic diagram of the growth process.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) the RHEED images of the AlGaN nanowires samples (<b>a</b>–<b>c</b>); (<b>d</b>–<b>f</b>) the SEM images of the AlGaN nanowires samples (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 3
<p>(<b>a</b>) STEM image of a single AlGaN nanowire, (<b>b</b>) EDX image along the growth direction, (<b>c</b>) HR-TEM images of the Qdisks region. (<b>d,e</b>) STEM images of the Qdisks region.</p>
Full article ">Figure 4
<p>(<b>a</b>) PL spectra of the AlGaN nanowires with different emission wavelengths, (<b>b</b>) TDPL spectra of the AlGaN nanowires emission at 305 nm, (<b>c</b>) IQE and wavelength change curves with temperature change, (<b>d</b>) the wavelength shifts fitting curve using the Varshni empirical formula.</p>
Full article ">
16 pages, 4733 KiB  
Article
Thermally Evaporated Copper Iodide Hole-Transporter for Stable CdS/CdTe Thin-Film Solar Cells
by Thuraisamykurukkal Thivakarasarma, Adikari Arachchige Isuru Lakmal, Buddhika Senarath Dassanayake, Dhayalan Velauthapillai and Punniamoorthy Ravirajan
Nanomaterials 2022, 12(14), 2507; https://doi.org/10.3390/nano12142507 - 21 Jul 2022
Cited by 2 | Viewed by 2217
Abstract
This study focuses on fabricating efficient CdS/CdTe thin-film solar cells with thermally evaporated cuprous iodide (CuI) as hole-transporting material (HTM) by replacing Cu back contact in conventional CdS/CdTe solar cells to avoid Cu diffusion. In this study, a simple thermal evaporation method was [...] Read more.
This study focuses on fabricating efficient CdS/CdTe thin-film solar cells with thermally evaporated cuprous iodide (CuI) as hole-transporting material (HTM) by replacing Cu back contact in conventional CdS/CdTe solar cells to avoid Cu diffusion. In this study, a simple thermal evaporation method was used for the CuI deposition. The current-voltage characteristic of devices with CuI films of thickness 5 nm to 30 nm was examined under illuminations of 100 mW/cm2 (1 sun) with an Air Mass (AM) of 1.5 filter. A CdS/CdTe solar cell device with thermally evaporated CuI/Au showed power conversion efficiency (PCE) of 6.92% with JSC, VOC, and FF of 21.98 mA/cm2, 0.64 V, and 0.49 under optimized fabrication conditions. Moreover, stability studies show that fabricated CdS/CdTe thin-film solar cells with CuI hole-transporters have better stability than CdS/CdTe thin-film solar cells with Cu/Au back contacts. The significant increase in FF and, hence, PCE, and the stability of CdS/CdTe solar cells with CuI, reveals that Cu diffusion could be avoided by replacing Cu with CuI, which provides good band alignment with CdTe, as confirmed by XPS. Such an electronic band structure alignment allows smooth hole transport from CdTe to CuI, which acts as an electron reflector. Hence, CuI is a promising alternative stable hole-transporter for CdS/CdTe thin-film solar cells that increases the PCE and stability. Full article
(This article belongs to the Section Solar Energy and Solar Cells)
Show Figures

Figure 1

Figure 1
<p>XRD pattern of thermally evaporated CuI film.</p>
Full article ">Figure 2
<p>3D and 2D AFM images of (<b>a</b>) CBD-CdS; (<b>b</b>) CdS/CSS-CdTe, and (<b>c</b>) CdS/CdTe/TE-CuI film.</p>
Full article ">Figure 3
<p>Schematic diagram of the four-point probe technique.</p>
Full article ">Figure 4
<p>Tauc plot and absorption spectrum (inset) of thermally evaporated CuI film.</p>
Full article ">Figure 5
<p>(<b>a</b>) Proposed energy alignment of FTO/CdS/CdTe/CuI/Au interface and (<b>b</b>) the quantitative electronic band alignment at the CSS-CdTe/TE-CuI/TE-Au interface obtained from XPS.</p>
Full article ">Figure 6
<p>(<b>a</b>) XPS survey spectrum of CdTe/CuI interface, (<b>b</b>) core level of the Cd 3d region, (<b>c</b>) core level of the Te 3d region, (<b>d</b>) core level of the Cu 2p region, (<b>e</b>) core level of the I 3d region, and (<b>f</b>) high-resolution spectra of the Cd 3d and Cu 2p regions in CdTe/CuI interface.</p>
Full article ">Figure 6 Cont.
<p>(<b>a</b>) XPS survey spectrum of CdTe/CuI interface, (<b>b</b>) core level of the Cd 3d region, (<b>c</b>) core level of the Te 3d region, (<b>d</b>) core level of the Cu 2p region, (<b>e</b>) core level of the I 3d region, and (<b>f</b>) high-resolution spectra of the Cd 3d and Cu 2p regions in CdTe/CuI interface.</p>
Full article ">Figure 7
<p>J–V characteristic curves of CBD-CdS/CSS-CdTe solar cells (without CdCl<sub>2</sub> treatment) with CuI with thicknesses of 5, 10, 15, 20, 25, and 30 nm.</p>
Full article ">Figure 8
<p>Photovoltaic parameters of (<b>a</b>) power conversion efficiency, (<b>b</b>) short circuit current density, (<b>c</b>) series resistance, (<b>d</b>) shunt resistance, and (<b>e</b>) fill factor of the CBD-CdS/CSS-CdTe devices with different CuI thicknesses.</p>
Full article ">Figure 9
<p>J–V curves of CBD-CdS/CSS-CdTe solar cells with a Cu/Au and CuI/Au back contact (<b>a</b>) under illumination of 100 mW/cm<sup>2</sup> with AM1.5 filter and (<b>b</b>) the semi-logarithmic J–V characteristics of CBD-CdS/CSS-CdTe solar cells with Cu/Au and CuI/Au back contact in dark.</p>
Full article ">Figure 10
<p>(<b>a</b>) J–V characteristic of CdS/CdTe/CuI/Au device fabricated with different CSS parameter under illumination of 100 mW/cm<sup>2</sup> with AM1.5 filter and (<b>b</b>) semi-logarithmic plot in dark.</p>
Full article ">Figure 11
<p>Energy band diagram of a generic FTO/n-CdS/p-CdTe/back-contact solar cell with the front interface, the depletion region, quasi-neutral region, and the back-contact interface.</p>
Full article ">
9 pages, 2629 KiB  
Article
Design and Simulation of Efficient SnS-Based Solar Cell Using Spiro-OMeTAD as Hole Transport Layer
by Pooja Tiwari, Maged F. Alotaibi, Yas Al-Hadeethi, Vaibhava Srivastava, Bassim Arkook, Sadanand, Pooja Lohia, Dilip Kumar Dwivedi, Ahmad Umar, Hassan Algadi and Sotirios Baskoutas
Nanomaterials 2022, 12(14), 2506; https://doi.org/10.3390/nano12142506 - 21 Jul 2022
Cited by 34 | Viewed by 2831
Abstract
In the present paper, the theoretical investigation of the device structure ITO/CeO2/SnS/Spiro-OMeTAD/Mo of SnS-based solar cell has been performed. The aim of this work is to examine how the Spiro-OMeTAD HTL affects the performance of SnS-based heterostructure solar cell. Using SCAPS-1D [...] Read more.
In the present paper, the theoretical investigation of the device structure ITO/CeO2/SnS/Spiro-OMeTAD/Mo of SnS-based solar cell has been performed. The aim of this work is to examine how the Spiro-OMeTAD HTL affects the performance of SnS-based heterostructure solar cell. Using SCAPS-1D simulation software, various parameters of SnS-based solar cell such as work function, series and shunt resistance and working temperature have been investigated. With the help of Spiro-OMeTAD, the suggested cell’s open-circuit voltage was increased to 344 mV. The use of Spiro-OMeTAD HTL in the SnS-based solar cell resulted in 14% efficiency increase, and the proposed heterojunction solar cell has 25.65% efficiency. The cell’s performance is determined by the carrier density and width of the CeO2 ETL (electron transport layer), SnS absorber layer and Spiro-OMeTAD HTL (hole transport layer). These data reveal that the Spiro-OMeTAD solar cells could have been a good HTL (hole transport layer) in regards to producing SnS-based heterojunction solar cell with high efficiency and reduced cost. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic layout of the proposed heterostructure solar cell and (<b>b</b>) the energy band diagram of the proposed heterostructure solar cell.</p>
Full article ">Figure 2
<p>(<b>a</b>) J-V curve and (<b>b</b>) quantum efficiency of a proposed solar cell.</p>
Full article ">Figure 3
<p>Effect on (<b>a</b>) J<sub>SC</sub>, (<b>b</b>), V<sub>oc</sub> (<b>c</b>) FF (<b>d</b>) and PCE with variation of work function (eV) of the proposed device.</p>
Full article ">Figure 4
<p>Effect on (<b>a</b>) J<sub>SC</sub>, (<b>b</b>)V<sub>OC</sub>, (<b>c</b>) FF and (<b>d</b>) PCE with variation of temperature of the proposed device.</p>
Full article ">Figure 5
<p>Effect on (<b>a</b>) J<sub>SC</sub>, (<b>b</b>) V<sub>OC,</sub> (<b>c</b>) FF and (<b>d</b>) PCE with variation of series resistance of the proposed device.</p>
Full article ">Figure 6
<p>Effect on (<b>a</b>) J<sub>SC</sub>, (<b>b</b>) V<sub>OC,</sub> (<b>c</b>) FF and (<b>d</b>) PCE with variation of shunt resistance of the proposed device.</p>
Full article ">
13 pages, 3550 KiB  
Article
MoS2-Decorated Graphene@porous Carbon Nanofiber Anodes via Centrifugal Spinning
by Elham Abdolrazzaghian, Jiadeng Zhu, Juran Kim and Meltem Yanilmaz
Nanomaterials 2022, 12(14), 2505; https://doi.org/10.3390/nano12142505 - 21 Jul 2022
Cited by 11 | Viewed by 2045
Abstract
Sodium-ion batteries (SIBs) are promising alternatives to lithium-ion batteries as green energy storage devices because of their similar working principles and the abundance of low-cost sodium resources. Nanostructured carbon materials are attracting great interest as high-performance anodes for SIBs. Herein, a simple and [...] Read more.
Sodium-ion batteries (SIBs) are promising alternatives to lithium-ion batteries as green energy storage devices because of their similar working principles and the abundance of low-cost sodium resources. Nanostructured carbon materials are attracting great interest as high-performance anodes for SIBs. Herein, a simple and fast technique to prepare carbon nanofibers (CNFs) is presented, and the effects of carbonization conditions on the morphology and electrochemical properties of CNF anodes in Li- and Na-ion batteries are investigated. Porous CNFs containing graphene were fabricated via centrifugal spinning, and MoS2 were decorated on graphene-included porous CNFs via hydrothermal synthesis. The effect of MoS2 on the morphology and the electrode performance was examined in detail. The results showed that the combination of centrifugal spinning, hydrothermal synthesis, and heat treatment is an efficient way to fabricate high-performance electrodes for rechargeable batteries. Furthermore, CNFs fabricated at a carbonization temperature of 800 °C delivered the highest capacity, and the addition of MoS2 improved the reversible capacity up to 860 mAh/g and 455 mAh/g for Li- and Na-ion batteries, respectively. A specific capacity of over 380 mAh/g was observed even at a high current density of 1 A/g. Centrifugal spinning and hydrothermal synthesis allowed for the fabrication of high-performance electrodes for sodium ion batteries. Full article
(This article belongs to the Special Issue Fabrication and Characterization of Nanostructured Carbon Electrodes)
Show Figures

Figure 1

Figure 1
<p>SEM images of (<b>A</b>) PAN nanofibers and carbon nanofibers carbonized at (<b>B</b>) 700 °C (<b>C</b>) 800 °C, and (<b>D</b>) 900 °C.</p>
Full article ">Figure 2
<p>XRD pattern (<b>a</b>) and RAMAN spectra (<b>b</b>) of carbon nanofibers carbonized at (<b>A</b>) 700 °C (<b>B</b>) 800 °C, and (<b>C</b>) 900 °C.</p>
Full article ">Figure 3
<p>First-cycle charge/discharge curves of carbon nanofibers carbonized at (<b>A</b>,<b>D</b>) 700 °C, (<b>B</b>,<b>E</b>) 800 °C, and (<b>C</b>,<b>F</b>) 900 °C in Li-ion cells and Na-ion batteries and cycling performance in Li-ion batteries (<b>G</b>) and Na-ion batteries (<b>H</b>).</p>
Full article ">Figure 4
<p>SEM images (<b>A</b>,<b>B</b>), of graphene@ porous carbon nanofiber and (<b>C</b>,<b>D</b>) MoS<sub>2</sub>-decorated graphene@ porous carbon nanofiber and TEM images of (<b>E</b>) graphene@ porous carbon nanofiber and (<b>F</b>) MoS<sub>2</sub>-decorated graphene@porous carbon nanofiber.</p>
Full article ">Figure 5
<p>XRD pattern (<b>a</b>) and Raman spectra, D and G bands of carbon (<b>b</b>) of graphene@ porous carbon nanofiber (<b>A</b>), MoS<sub>2</sub>-decorated graphene@ porous carbon nanofiber (<b>B</b>), and TGA curve of MoS<sub>2</sub>-decorated on graphene@PCNFs (<b>c</b>).</p>
Full article ">Figure 6
<p>First-cycle discharge/charge curves and cycling performance of graphene@ porous carbon nanofiber electrodes in Li-ion (<b>A</b>,<b>C</b>) and Na-ion batteries (<b>B</b>,<b>D</b>).</p>
Full article ">Figure 7
<p>First-cycle discharge/charge curves and cycling performance of MoS<sub>2</sub>-decorated graphene@porous carbon nanofiber electrodes in Li-ion (<b>A</b>,<b>C</b>) and Na-ion batteries (<b>B</b>–<b>E</b>).</p>
Full article ">
14 pages, 3346 KiB  
Article
Tailoring of AlAs/InAs/GaAs QDs Nanostructures via Capping Growth Rate
by Nazaret Ruiz, Daniel Fernandez, Esperanza Luna, Lazar Stanojević, Teresa Ben, Sara Flores, Verónica Braza, Alejandro Gallego-Carro, Guillermo Bárcena-González, Andres Yañez, José María Ulloa and David González
Nanomaterials 2022, 12(14), 2504; https://doi.org/10.3390/nano12142504 - 21 Jul 2022
Cited by 1 | Viewed by 1630
Abstract
The use of thin AlA capping layers (CLs) on InAs quantum dots (QDs) has recently received considerable attention due to improved photovoltaic performance in QD solar cells. However, there is little data on the structural changes that occur during capping and their relation [...] Read more.
The use of thin AlA capping layers (CLs) on InAs quantum dots (QDs) has recently received considerable attention due to improved photovoltaic performance in QD solar cells. However, there is little data on the structural changes that occur during capping and their relation to different growth conditions. In this work, we studied the effect of AlA capping growth rate (CGR) on the structural features of InAs QDs in terms of shape, size, density, and average content. As will be shown, there are notable differences in the characteristics of the QDs upon changing CGR. The Al distribution analysis in the CL around the QDs was revealed to be the key. On the one hand, for the lowest CGR, Al has a homogeneous distribution over the entire surface, but there is a large thickening of the CL on the sides of the QD. As a result, the QDs are lower, lenticular in shape, but richer in In. On the other hand, for the higher CGRs, Al accumulates preferentially around the QD but with a more uniform thickness, resulting in taller QDs, which progressively adopt a truncated pyramidal shape. Surprisingly, intermediate CGRs do not improve either of these behaviors, resulting in less enriched QDs. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the sample structure. The growth rate of the AlAs CLs increased in each layer, 0.25, 0.5, 0.75 and 1 ML/s, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) LAADF images of the layers CL1 (top), CL0.25 (middle) and CL0 (bottom). Lens and pyramidal-shaped dots are marked with L and P, respectively. The Al-rich CLs appear darker. (<b>b</b>) Zoom of a truncated pyramidal (left) and lenticular (right) QD capped with AlAs. (<b>c</b>) The proportion of lenticular and pyramidal QDs in the QD population for the different QDs layers.</p>
Full article ">Figure 3
<p>Histograms of (<b>a</b>) QD height and (<b>b</b>) base diameter distributions for lens and truncated pyramidal geometries of the different QD layers. The differences in the distribution between the two geometries become smaller as the CGR increased. Arrows show the position of the mean values of lenticular (red) and pyramidal (black) QDs. (<b>c</b>) The volume of QDs for lenticular and truncated pyramidal-shaped QDs versus CGR. Both geometries show similar average volume for the fastest CGR. (<b>d</b>) Average height and diameter of QDs versus CGR. The sample without CL of AlAs (CL0) is not connected by lines.</p>
Full article ">Figure 4
<p>(<b>a</b>) Elemental EDX maps of In and Al for the different QD layers. (<b>b</b>) Average compositional profiles of In (red) and Al (green) along the growth direction in the CL/WL regions of the QD layers CL0.75 and CL0. All the compositional profiles of the QD layers with CL of AlAs in the regions between dots are remarkably similar. The inset corresponds to a zoom of the gray region.</p>
Full article ">Figure 5
<p>(<b>a</b>) Average QD volume (left axis) and superficial density (right axis) versus CGR (<b>b</b>) Average In content obtained for the different layers. Better protection of dots is seen for extreme AlAs CGR.</p>
Full article ">Figure 6
<p>In (red) and Al (green) EDX maps of representative lens-shaped QDs (the top two rows) for the QD layers with CL of AlAs. We have also included examples of truncated pyramidal-shaped QDs (the bottom two rows) when the proportion of pyramids is significant (CL0.75 and CL1 layers).</p>
Full article ">Figure 7
<p>(<b>a</b>) Scheme of a QD capped by an AlAs layer. L1 to L4 are directions parallel to the growth plane where the compositional profiles are taken. P1 to P4 are the positions for punctual measurements. (<b>b</b>) Experimental compositional profiles along the line L3 for a QD in the CL1 layer. (<b>c</b>) Plan view of a slide of the QD capped with AlAs and its correspondence with the theoretical composition profiles for In and Al.</p>
Full article ">Figure 8
<p>(<b>a</b>) Average contours of the QD and CL for layers CL0.25 and CL1. Lateral thicknesses of the CL at the base and close to the apex are named <span class="html-italic">t</span>1 and <span class="html-italic">t</span>3, respectively. (<b>b</b>) Lateral thicknesses <span class="html-italic">t1</span> and <span class="html-italic">t3</span> of the CL versus the CGR.</p>
Full article ">Figure 9
<p>Al contents in the CL for lenticular (<b>a</b>) and truncated pyramid QDs (<b>b</b>) in the positions marked in <a href="#nanomaterials-12-02504-f007" class="html-fig">Figure 7</a>. (P1) corresponds to regions of the WL away from the QD, (P2) at the QD edge, (P3) close to the QD top and (P4) above the QD.</p>
Full article ">Figure 10
<p>Schematic representation of the CGR effects in the GaAs/InAs/AlAs QD system for a (<b>a</b>) QD capped with GaAs (CL0), (<b>b</b>) covered by AlAs at 0.25 ML/s (CL0.25), (<b>c</b>) 0.5–0.75 ML/s and (<b>d</b>) 1 ML/s (CL1). “L” and “P” at the bottom right in each picture designate the lens and truncated pyramidal geometries. Average compositions for the QDs and the CL are indicated.</p>
Full article ">
18 pages, 4224 KiB  
Article
Evolution of the Electronic and Optical Properties of Meta-Stable Allotropic Forms of 2D Tellurium for Increasing Number of Layers
by Simone Grillo, Olivia Pulci and Ivan Marri
Nanomaterials 2022, 12(14), 2503; https://doi.org/10.3390/nano12142503 - 21 Jul 2022
Cited by 5 | Viewed by 2490
Abstract
In this work, ab initio Density Functional Theory calculations are performed to investigate the evolution of the electronic and optical properties of 2D Tellurium—called Tellurene—for three different allotropic forms (α-, β- and γ-phase), as a function of the number [...] Read more.
In this work, ab initio Density Functional Theory calculations are performed to investigate the evolution of the electronic and optical properties of 2D Tellurium—called Tellurene—for three different allotropic forms (α-, β- and γ-phase), as a function of the number of layers. We estimate the exciton binding energies and radii of the studied systems, using a 2D analytical model. Our results point out that these quantities are strongly dependent on the allotropic form, as well as on the number of layers. Remarkably, we show that the adopted method is suitable for reliably predicting, also in the case of Tellurene, the exciton binding energy, without the need of computationally demanding calculations, possibly suggesting interesting insights into the features of the system. Finally, we inspect the nature of the mechanisms ruling the interaction of neighbouring Tellurium atoms helical chains (characteristic of the bulk and α-phase crystal structures). We show that the interaction between helical chains is strong and cannot be explained by solely considering the van der Waals interaction. Full article
Show Figures

Figure 1

Figure 1
<p>Perspective and side views of the crystal structures of bilayer (2L) <math display="inline"><semantics> <mi>α</mi> </semantics></math>-, <math display="inline"><semantics> <mi>β</mi> </semantics></math>- and <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-phase of Tellurene.</p>
Full article ">Figure 2
<p>Total energy per atom (rescaled with respect to an isolated Te atom), for increasing number of layers, of the three studied phases. The <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-phase is the most stable in the ML configuration, while the <math display="inline"><semantics> <mi>α</mi> </semantics></math>-phase is preferred for larger layer thicknesses. Note that the ML <math display="inline"><semantics> <mi>α</mi> </semantics></math>-phase is unstable.</p>
Full article ">Figure 3
<p>Electronic bandstructures, obtained by using a PBE functional, with the inclusion of SOC, for ML <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-Te (<b>a</b>) and <math display="inline"><semantics> <mi>β</mi> </semantics></math>-Te (<b>b</b>). Energy rescaled with respect to the VBM. 2D hexagonal (<b>c</b>, <b>left</b>) and orthorhombic (<b>c</b>, <b>right</b>) BZ, with the high symmetry points used in electronic bandstructure calculations.</p>
Full article ">Figure 4
<p>Electronic bandstructures, obtained by using a PBE functional, with the inclusion of SOC, for 2L <math display="inline"><semantics> <mi>β</mi> </semantics></math>-Te (<b>a</b>), 3L <math display="inline"><semantics> <mi>β</mi> </semantics></math>-Te (<b>b</b>) and 4L <math display="inline"><semantics> <mi>β</mi> </semantics></math>-Te (<b>c</b>). Energy rescaled with respect to the VBM. High-symmetry points related to <a href="#nanomaterials-12-02503-f003" class="html-fig">Figure 3</a> (orthorhombic BZ).</p>
Full article ">Figure 5
<p>Electronic bandstructures, obtained by using a PBE functional, with the inclusion of SOC, for 2L <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-Te (<b>a</b>), 3L <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-Te (<b>b</b>) and 4L <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-Te (<b>c</b>). Energy rescaled with respect to the VBM (and Fermi energy for 4L). High-symmetry points related to <a href="#nanomaterials-12-02503-f003" class="html-fig">Figure 3</a> (hexagonal BZ).</p>
Full article ">Figure 6
<p>Electronic bandstructures, obtained by using a PBE functional, with the inclusion of SOC, for 2L <math display="inline"><semantics> <mi>α</mi> </semantics></math>-Te (<b>a</b>), 3L <math display="inline"><semantics> <mi>α</mi> </semantics></math>-Te (<b>b</b>) and 4L <math display="inline"><semantics> <mi>α</mi> </semantics></math>-Te (<b>c</b>). Energy rescaled with respect to the VBM. High-symmetry points related to <a href="#nanomaterials-12-02503-f003" class="html-fig">Figure 3</a> (orthorhombic BZ). Note that lower values of the gaps were found out of high-symmetry directions and they are reported in <a href="#nanomaterials-12-02503-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 7
<p>In-plane optical absorbance of (<b>a</b>) ML <math display="inline"><semantics> <mi>β</mi> </semantics></math>- and (<b>b</b>) <math display="inline"><semantics> <mi>γ</mi> </semantics></math>-Te, with the inclusion of SOC. Absorption energy threshold estimated values of 1.02 eV and 0.54 eV, respectively.</p>
Full article ">Figure 8
<p>In-plane optical absorbance comparison between 2L, 3L and 4L <math display="inline"><semantics> <mi>γ</mi> </semantics></math>- (<b>a</b>), <math display="inline"><semantics> <mi>β</mi> </semantics></math>- (<b>b</b>) and <math display="inline"><semantics> <mi>α</mi> </semantics></math>-Te (<b>c</b>), with the inclusion of SOC. Overall, the absorption energy threshold decreases for increasing number of layers (see <a href="#nanomaterials-12-02503-t002" class="html-table">Table 2</a>).</p>
Full article ">Figure 9
<p>Numerical solutions of the 2D exciton model, for <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>B</mi> </msub> <mo>/</mo> <msub> <mi>R</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> (red) and <math display="inline"><semantics> <mrow> <msub> <mi>r</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> <mo>/</mo> <msub> <mi>a</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </mrow> </semantics></math> (blue), as expressed by Equation (<a href="#FD5-nanomaterials-12-02503" class="html-disp-formula">5</a>), showing the two limits discussed. <math display="inline"><semantics> <msub> <mi>E</mi> <mi>B</mi> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>r</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> </semantics></math> are the exciton binding energy and radius, respectively; <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>R</mi> <mi>H</mi> </msub> <mfrac> <mi>μ</mi> <mi>m</mi> </mfrac> </mrow> </semantics></math> is the 3D hydrogenoid Rydberg <math display="inline"><semantics> <msub> <mi>R</mi> <mi>H</mi> </msub> </semantics></math>, renormalised by the ratio between the effective reduced mass <math display="inline"><semantics> <mi>μ</mi> </semantics></math> and the free electron mass <span class="html-italic">m</span>; <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mrow> <mi>e</mi> <mi>x</mi> </mrow> </msub> <mo>=</mo> <msub> <mi>a</mi> <mi>B</mi> </msub> <mfrac> <mi>m</mi> <mi>μ</mi> </mfrac> </mrow> </semantics></math> is the Bohr radius <math display="inline"><semantics> <msub> <mi>a</mi> <mi>B</mi> </msub> </semantics></math>, renormalised by the free electron mass and the effective reduced mass ratio. Results for Te are all from this work. Credits to [<a href="#B37-nanomaterials-12-02503" class="html-bibr">37</a>,<a href="#B38-nanomaterials-12-02503" class="html-bibr">38</a>,<a href="#B39-nanomaterials-12-02503" class="html-bibr">39</a>] for the other results reported (see top left inset): InN (green), GaN and Graphane (grey), BN (orange), AlN (cyan), Plumbene:H (magenta).</p>
Full article ">Figure 10
<p>Pictorial scheme of a general atom-atom bonding length distribution, as described by Alvarez [<a href="#B54-nanomaterials-12-02503" class="html-bibr">54</a>].</p>
Full article ">Figure 11
<p>Interchain binding energy as a function of the interchain distance, with and without the inclusion of the Grimme’s DFT-D2 vdW correction. In both case, the chains possess the same fixed geometry. Red dots correspond to the equilibrium interchain separation of minimum energy for the two cases. Energy rescaled with respect to the relative isolated chains (with and without vdW).</p>
Full article ">Figure 12
<p>Electron localization function (ELF) for two Te helical chains. Vertical 2D plot cutting through a Te−Te bond axis along a chain (<b>a</b>) and between the chains (<b>b</b>).</p>
Full article ">
13 pages, 1749 KiB  
Article
Isolation and Characterization of Cell Envelope Fragments Comprising Archaeal S-Layer Proteins
by Kevin Pfeifer, Eva-Kathrin Ehmoser, Simon K.-M. R. Rittmann, Christa Schleper, Dietmar Pum, Uwe B. Sleytr and Bernhard Schuster
Nanomaterials 2022, 12(14), 2502; https://doi.org/10.3390/nano12142502 - 21 Jul 2022
Cited by 2 | Viewed by 2002
Abstract
The outermost component of cell envelopes of most bacteria and almost all archaea comprise a protein lattice, which is termed Surface (S-)layer. The S-layer lattice constitutes a highly porous structure with regularly arranged pores in the nm-range. Some archaea thrive in extreme milieus, [...] Read more.
The outermost component of cell envelopes of most bacteria and almost all archaea comprise a protein lattice, which is termed Surface (S-)layer. The S-layer lattice constitutes a highly porous structure with regularly arranged pores in the nm-range. Some archaea thrive in extreme milieus, thus producing highly stable S-layer protein lattices that aid in protecting the organisms. In the present study, fragments of the cell envelope from the hyperthermophilic acidophilic archaeon Saccharolobus solfataricus P2 (SSO) have been isolated by two different methods and characterized. The organization of the fragments and the molecular sieving properties have been elucidated by transmission electron microscopy and by determining the retention efficiency of proteins varying in size, respectively. The porosity of the archaeal S-layer fragments was determined to be 45%. S-layer fragments of SSO showed a retention efficiency of up to 100% for proteins having a molecular mass of ≥ 66 kDa. Moreover, the extraction costs for SSO fragments have been reduced by more than 80% compared to conventional methods, which makes the use of these archaeal S-layer material economically attractive. Full article
(This article belongs to the Section Nanocomposite Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) <span class="html-italic">Saccharolobus solfataricus</span> P2 (SSO) ghost prepared using the extraction with DNase I (DEG). (<b>B</b>,<b>C</b>) SSO ghosts prepared using the extraction method with sonication (SEG). Scale Bars: A = 100 nm; B = 500 nm; C = 250 nm.</p>
Full article ">Figure 2
<p>SDS-PAGE (4–20% Bis-Tris, MES) S1–S3: 20 µL of <span class="html-italic">Saccharolobus solfataricus</span> P2 (SSO) sacculi using the extraction method with sonication (SEG). D1–D3: 20 µL of <span class="html-italic">SSO</span> sacculi using the established extraction method with DNase I (DEG). The first lane corresponds to molecular mass standard with ladder units in kDa.</p>
Full article ">Figure 3
<p>(<b>A</b>) Transmission electron micrograph of <span class="html-italic">Saccharolobus solfataricus</span> P2 (SSO) fragments after sonication. Bar, 1000 nm; (<b>B</b>) computer image reconstruction (2D projection) of the p3-ordered hexagonal S-layer lattice from SSO. Shown are a unit cell and the corresponding three-fold axis of rotation (symmetry operators) in blue. The protein is light, the pores are dark. In addition, a possible configuration of proteins belonging to a unit cell is shown by strokes drawn by hand in orange. Bar, 20 nm; (<b>C</b>) scanning electron micrograph of SSO fragments deposited on microfilter 1 (100,000× magnification). Bar, 500 nm.</p>
Full article ">Figure 4
<p>(<b>A</b>) Average efficiency of <span class="html-italic">Saccharolobus solfataricus</span> P2 fragments deposited on microfilter 1 (SSOMF1) across all experiments (<span class="html-italic">n</span> ≥ 6); (<b>B</b>) retention efficiency of SSOMF1 where Orange = filtration sequence largest to smallest protein (<span class="html-italic">n</span> ≥ 3) and Blue = filtration sequence smallest to largest protein (<span class="html-italic">n</span> ≥ 3); Legend for A and B: ▬ = myoglobin (17 kDa); ■ = carbonic anhydrase (31 kDa); ♦ = horseradish peroxidase (44 kDa); ▲ = bovine serum albumin (66 kDa); <b>x</b> = amyloglucosidase (97 kDa) and ● = γ-globulin (125 kDa).</p>
Full article ">Figure 5
<p>Retention efficiency of <span class="html-italic">Saccharolobus</span> <span class="html-italic">solfataricus</span> P2 fragments deposited on microfilter 2 (SSOMF2) (orange); <span class="html-italic">B. stearothermophilus</span> pV72 (yellow) and <span class="html-italic">C. thermohydrosulfuricum</span> L111-69 (gray). The proteins with increasing molecular mass are myoglobin (17 kDa), carbonic anhydrase (31 kDa), ovalbumin (43 kDa; for bacterial SLP material), horseradish peroxidase (44 kDa; for SSOMF2), bovine serum albumin (66 kDa), and amyloglucosidase (97 kDa) (n ≥ 4). Data for the retention curve of <span class="html-italic">B. stearothermophilus</span> pV72 and <span class="html-italic">C. thermohydrosulfuricum</span> L111-69 are taken from Ref. [<a href="#B25-nanomaterials-12-02502" class="html-bibr">25</a>].</p>
Full article ">
35 pages, 4514 KiB  
Review
Semiconductor Quantum Dots as Target Analytes: Properties, Surface Chemistry and Detection
by Jesús Sanmartín-Matalobos, Pilar Bermejo-Barrera, Manuel Aboal-Somoza, Matilde Fondo, Ana M. García-Deibe, Julio Corredoira-Vázquez and Yeneva Alves-Iglesias
Nanomaterials 2022, 12(14), 2501; https://doi.org/10.3390/nano12142501 - 21 Jul 2022
Cited by 19 | Viewed by 4622
Abstract
Since the discovery of Quantum Dots (QDs) by Alexey I. Ekimov in 1981, the interest of researchers in that particular type of nanomaterials (NMs) with unique optical and electrical properties has been increasing year by year. Thus, since 2009, the number of scientific [...] Read more.
Since the discovery of Quantum Dots (QDs) by Alexey I. Ekimov in 1981, the interest of researchers in that particular type of nanomaterials (NMs) with unique optical and electrical properties has been increasing year by year. Thus, since 2009, the number of scientific articles published on this topic has not been less than a thousand a year. The increasing use of QDs due to their biomedical, pharmaceutical, biological, photovoltaics or computing applications, as well as many other high-tech uses such as for displays and solid-state lighting (SSL), has given rise to a considerable number of studies about its potential toxicity. However, there are a really low number of reported studies on the detection and quantification of QDs, and these include ICP–MS and electrochemical analysis, which are the most common quantification techniques employed for this purpose. The knowledge of chemical phenomena occurring on the surface of QDs is crucial for understanding the interactions of QDs with species dissolved in the dispersion medium, while it paves the way for a widespread use of chemosensors to facilitate its detection. Keeping in mind both human health and environmental risks of QDs as well as the scarcity of analytical techniques and methodological approaches for their detection, the adaptation of existing techniques and methods used with other NMs appears necessary. In order to provide a multidisciplinary perspective on QD detection, this review focused on three interrelated key aspects of QDs: properties, surface chemistry and detection. Full article
(This article belongs to the Special Issue Nanoparticle Analysis, Toxicity and Environmental Impact)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Formation of the exciton from the electron-hole pair and subsequent electron-hole recombination.</p>
Full article ">Figure 2
<p>Schematic representation of the three main types of CS QDs: Type I, Inverse type I and Type II, of which CdTe/ZnS, CdS/CdSe and CdTe/CdSe, as well as CdS/ZnSe, are examples. Band gap energies (<span class="html-italic">E</span><sub>g</sub>) for the semiconductor materials used as examples are indicated. Electron-hole recombination is indicated by a wavy arrow.</p>
Full article ">Figure 3
<p>Schematic diagram to illustrate the bonding forms for main adsorption modes of dissolved ligands onto the surface of CdSe QDs (Cd: grey balls; Se: brown balls). Left top: CdSe-octylamine has been used here as an illustrative example of an H-bonding interaction. Right top: The interaction between CdSe and a catechol-based oleylamine shows a couple of modes of chelation. Right-bottom: The desorption of oleic acid from coated QDs and the subsequent absorption of 4-mercaptobenzoic acid illustrates the ligand exchange mode. Left-bottom: The interaction between thioglycolate-capped QDs and Ca<sup>2+</sup> shows cation bridging mode among particles.</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectrum (in D<sub>2</sub>O) of cysteine-capped CdSe QDs (<b>top</b>). The spectrum of <span class="html-small-caps">L</span>-cysteine (<b>bottom</b>) has been included for comparison.</p>
Full article ">Figure 5
<p>ATR–FTIR spectrum cysteine-capped CdSe QDs (<b>top</b>). The spectrum of <span class="html-small-caps">L</span>-cysteine (<b>bottom</b>) has been included for comparison purposes.</p>
Full article ">Figure 6
<p>UV–Vis (<b>left</b>) and fluorescence (<b>right</b>) spectra of the cysteine-capped CdSe QDs in ethanol (λ<sub>ex</sub> = 400 nm; λ<sub>em</sub> = 632 nm).</p>
Full article ">Figure 7
<p>(<b>a</b>) EDX spectrum of CdSe–Cys QDs; (<b>b</b>): SEM micrograph with EDX mapping pattern of the elements Cd, Se, C, S, N, Na and O from an anionic cysteine-capped CdSe QDs sample.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) EDX spectrum of CdSe–Cys QDs; (<b>b</b>): SEM micrograph with EDX mapping pattern of the elements Cd, Se, C, S, N, Na and O from an anionic cysteine-capped CdSe QDs sample.</p>
Full article ">Figure 8
<p>Sector graph of analytical techniques for detection/determination of QDs.</p>
Full article ">
3 pages, 204 KiB  
Editorial
Functional Biodegradable Nanocomposites
by Agueda Sonseca, Coro Echeverría and Daniel López
Nanomaterials 2022, 12(14), 2500; https://doi.org/10.3390/nano12142500 - 21 Jul 2022
Viewed by 1233
Abstract
Over 367 million tons of plastics are produced annually worldwide, and the growth of plastic pollution has become a global concern [...] Full article
(This article belongs to the Special Issue Functional Biodegradable Nanocomposites)
11 pages, 1970 KiB  
Article
A Self-Assembly of Single Layer of Co Nanorods to Reveal the Magnetostatic Interaction Mechanism
by Hongyu Du, Min Zhang, Ke Yang, Baohe Li and Zhenhui Ma
Nanomaterials 2022, 12(14), 2499; https://doi.org/10.3390/nano12142499 - 21 Jul 2022
Cited by 2 | Viewed by 1459
Abstract
In this work, we report a self-assembly method to fabricate a single layer of Co nanorods to study their magnetostatic interaction behavior. The Co nanorods with cambered and flat tips were synthesized by using a solvothermal route and an alcohol–thermal method, respectively. Both [...] Read more.
In this work, we report a self-assembly method to fabricate a single layer of Co nanorods to study their magnetostatic interaction behavior. The Co nanorods with cambered and flat tips were synthesized by using a solvothermal route and an alcohol–thermal method, respectively. Both of them represent hard magnetic features. Co nanorods with cambered tips have an average diameter of 10 nm and length of 100 nm with coercivity of 6.4 kOe, and flat-tip nanorods with a 30 nm diameter and 100 nm length exhibit a coercivity of 4.9 kOe. They are further assembled on the surface of water in assistance of surfactants. The results demonstrate that the assembly type is dependent on the magnetic induction lines direction. For Co nanorods with flat tips, most of magnetic induction lines are parallel to the length direction, leading to an assembly that is tip to tip. For Co nanorods with cambered tips, they are prone to holding together side by side for their random magnetic induction lines. Under an applied field, the Co nanorods with flat tips can be further aligned into a single layer of Co nanorods. Our work gives a possible mechanism for the magnetic interaction of Co nanorods and provides a method to study their magnetic behavior. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Co nanorods with cambered tips by solvothermal route: (<b>A</b>) XRD pattern compared with standard hcp-Co (JCPDS No. 01-1278); (<b>B</b>,<b>C</b>) TEM images with different magnification; (<b>D</b>) HRTEM image.</p>
Full article ">Figure 2
<p>(<b>A</b>–<b>C</b>) The TEM images of self-assembly Co nanorods with cambered tips by solvothermal route; (<b>D</b>) Magnetic hysteresis loop of Co nanorods at room temperature.</p>
Full article ">Figure 3
<p>The Co nanorods with flat tips by alcohol–thermal method. (<b>A</b>) XRD pattern compared with standard hcp-Co (JCPDS No. 01-1278); (<b>B</b>,<b>C</b>) TEM images of Co nanorods with 100 r/min stirring rate; (<b>D</b>) TEM image of Co nanorods with 50 r/min stirring rate.</p>
Full article ">Figure 4
<p>Magnetic hysteresis loop of Co nanorods with flat tips at room temperature.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) The TEM images of self-assembly Co nanorods with flat tips, without applied field; (<b>C</b>) XRD pattern of aligned Co nanorods under magnetic field compared with standard hcp-Co (JCPDS No. 01-1278); (<b>D</b>) TEM image of aligned Co nanorods under magnetic field.</p>
Full article ">Figure 6
<p>The scheme of magnetostatic interaction mechanism: (<b>A</b>) Co nanorods with flat tips and (<b>B</b>) Co nanorods with cambered tips.</p>
Full article ">
8 pages, 2482 KiB  
Article
Boosting the Humidity Performances of Na0.5BixTiO3 by Tuning Bi Content
by Xiaoqi Xuan, Li Li, Tiantian Li, Jingsong Wang, Yi Yu and Chunchang Wang
Nanomaterials 2022, 12(14), 2498; https://doi.org/10.3390/nano12142498 - 21 Jul 2022
Viewed by 1285
Abstract
In the field of humidity sensors, a major challenge is how to improve the sensing performance of existing materials. Based on our previous work on Na0.5Bi0.5TiO3, a facile strategy of tuning the Bi content in the material [...] Read more.
In the field of humidity sensors, a major challenge is how to improve the sensing performance of existing materials. Based on our previous work on Na0.5Bi0.5TiO3, a facile strategy of tuning the Bi content in the material was proposed to improve its sensing performance. Na0.5BixTiO3 (x = 0.3, 0.35, 0.4, 0.45) nanocomposites were synthesized by a hydrothermal method. Humidity sensing properties of these nanocomposites were investigated in the relative humidity range of 11% to 95%. Our results show that, compared to the sensor based on nominally pure sample (Na0.5Bi0.5TiO3), the sensor based on Na0.5Bi0.35TiO3 exhibits boosted sensing performance of excellent linear humidity response in the humidity range of 11–75% relative humidity, lower hysteresis value, and faster response/recovery time. The improvement of the sensing performance was argued to be the reason that the proper reduction in Bi content leads to a minimum value of oxygen-vacancy concentrations, thereby weakening the chemical adsorption but enhancing the physical adsorption. These results indicate that the proper underdose of the Bi content in Na0.5Bi0.5TiO3 can greatly boost the sensing performance. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns and (<b>b</b>–<b>f</b>) SEM images of the Na<sub>0.5</sub>BixTiO<sub>3</sub> powders.</p>
Full article ">Figure 2
<p>(<b>a</b>) Humidity sensing properties of Na<sub>0.5</sub>Bi<sub>x</sub>TiO<sub>3</sub>-based sensor recorded with 200 Hz. The inset graph shows the linear fitting of the impedance data over the humidity range of 11–75% RH for the Na<sub>0.5</sub>Bi<sub>0.35</sub>TiO<sub>3</sub>-based sensor. (<b>b</b>) Hysteresis behavior of the Na<sub>0.5</sub>Bi<sub>0.35</sub>TiO<sub>3</sub>—based sensor.</p>
Full article ">Figure 3
<p>Humidity sensing properties of Na<sub>0.5</sub>Bi<sub>0.35</sub>TiO<sub>3</sub>-based sensor recorded with 200 Hz. (<b>a</b>) Repeatability curve, (<b>b</b>) response and recovery curve, and (<b>c</b>) long-term stability curves under different humidity levels.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) The complex impedance diagrams of Na<sub>0.5</sub>Bi<sub>x</sub>TiO<sub>3</sub>.</p>
Full article ">Figure 5
<p>High-resolution XPS spectra of O 1 s of the Na<sub>0.5</sub>Bi<sub>0.35</sub>TiO<sub>3</sub> (<b>a</b>) and Na<sub>0.5</sub>Bi<sub>0.5</sub>TiO<sub>3</sub> (<b>b</b>).</p>
Full article ">
21 pages, 5922 KiB  
Article
All-Solution Processed Single-Layer WOLEDs Using [Pt(salicylidenes)] as Guests in a PFO Matrix
by José Carlos Germino, Luís Gustavo Teixeira Alves Duarte, Rodrigo Araújo Mendes, Marcelo Meira Faleiros, Andreia de Morais, Jilian Nei de Freitas, Luiz Pereira and Teresa Dib Zambon Atvars
Nanomaterials 2022, 12(14), 2497; https://doi.org/10.3390/nano12142497 - 20 Jul 2022
Cited by 3 | Viewed by 2142
Abstract
Herein, we report the synthesis and characterization of two Pt(II) coordination compounds, the new platinum(II)[N,N′-bis(salicylidene)-3,4-diaminobenzophenone)] ([Pt(sal-3,4-ben)]) and the already well-known platinum(II)[N,N′-bis(salicylidene)-o-phenylenediamine] ([Pt(salophen)]), along with their application as guests in a poly [9,9-dioctylfluorenyl-2,7-diyl] (PFO) conjugated polymer [...] Read more.
Herein, we report the synthesis and characterization of two Pt(II) coordination compounds, the new platinum(II)[N,N′-bis(salicylidene)-3,4-diaminobenzophenone)] ([Pt(sal-3,4-ben)]) and the already well-known platinum(II)[N,N′-bis(salicylidene)-o-phenylenediamine] ([Pt(salophen)]), along with their application as guests in a poly [9,9-dioctylfluorenyl-2,7-diyl] (PFO) conjugated polymer in all-solution processed single-layer white organic light-emitting diodes. Completely different performances were achieved: 2.2% and 15.3% of external quantum efficiencies; 2.8 cd A−1 and 12.1 cd A−1 of current efficiencies; and 3103 cd m−2 and 6224 cd m−2 of luminance for the [Pt(salophen)] and [Pt(sal-3,4-ben)] complexes, respectively. The Commission Internationale de l’Eclairage (CIE 1931) chromaticity color coordinates are (0.33, 0.33) for both 0.1% mol/mol Pt(II):PFO composites at between approximately 3.2 and 8 V. The optoelectronic properties of doped and neat PFO films have been investigated, using steady-state and time-resolved photoluminescence. Theoretical calculations at the level of relativistic density functional theory explained these results, based on the presence of the Pt(II) central ion’s phosphorescence emission, considering spin-orbit coupling relationships. The overall results are explained, taking into account the active layer morphological properties, along with the device’s electric balance and the emitter’s efficiencies, according to deep-trap space-charge models. Considering the very simple structure of the device and the ease of synthesis of such compounds, the developed framework can offer a good trade-off for solution-deposited white organic light-emitting diodes (WOLEDs), with further applications in the field of lighting and signage. Full article
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structures of PFO (<b>a</b>), [Pt(salophen)] (<b>b</b>) and [Pt(sal-3,4-ben)] (<b>c</b>) organic semiconducting materials.</p>
Full article ">Figure 2
<p>Molecular crystal structures of the [Pt(salophen)] and [Pt(sal-3,4-ben)] coordination compound ellipsoids, with 50% probability, where: hydrogen is the white balls; carbon is the gray ellipsoids; nitrogen is the blue ellipsoids; oxygen is the red ellipsoids; and platinum is the dark-blue ellipsoids.</p>
Full article ">Figure 3
<p>(<b>a</b>) NTOs hole/particle pair densities for the first triplet excited states of [Pt(salophen)] and [Pt(sal-3,4-ben)] in a vacuum at the PBE0/(def2-TZVP(C,N,O,H) and SARC-ZORA-def2-TZVP-Pt(II) atom) level (green and red colors represent the different density parities); (<b>b</b>) experimental OLED energy level diagram (inside the PFO energy diagram: blue represents [Pt(salophen)] and red represents [Pt(sal-3,4-ben)]).</p>
Full article ">Figure 4
<p>Normalized PL (λ<sub>exc</sub> = 375 nm) spectra of the PFO composites. (<b>a</b>) [Pt(salophen)] and (<b>b</b>) [Pt(sal-3,4-ben)], with 0%, 0.1%, 0.5% and 1% mol/mol of complexes.</p>
Full article ">Figure 5
<p>Electroluminescence spectra for [Pt(salophen)] and [Pt(sal-3,4-ben)] OLEDs with different concentrations, at maximum EQE. At the bottom, CIE1931 chromaticity coordinates the dependence on the bias of PFO:[Pt(II)]-based OLEDs for concentrations of 0.1, 0.5, and 1.0% of the Pt(II) complexes (arrows indicate the applied voltage increase and solid-lines represents the WOLEDs color change).</p>
Full article ">Figure 6
<p>Optical-electronic properties of the diodes ITO|PEDOT:PSS/PVK/PFO:[Pt(salicylidene)s]/Ca|Al: current density and brightness vs. the voltage data of [Pt(salophen)] (<b>a</b>,<b>c</b>) and [Pt(sal-3,4-ben)] (<b>b</b>,<b>d</b>)-based OLEDs.</p>
Full article ">Figure 7
<p>OLEDs figures of merit vs. brightness plots for PFO:[Pt(salophen)s] (<b>a</b>) and PFO:[Pt(sal-3,4-ben)] (<b>b</b>).</p>
Full article ">Figure 8
<p>Density of the trap states, as determined by the VTFL data for both OLEDs’ emissive layers.</p>
Full article ">Figure 9
<p>AFM topography images (45 × 45 μm<sup>2</sup>) of the PFO:Pt(salicylidenes) emissive layers, on top of the ITO|PEDOT:PSS|PVK layers: [Pt(salophen)] (<b>a</b>,<b>b</b>) and [Pt(sal-3,4-ben)] (<b>c</b>,<b>d</b>) at 0.1% and 0.5%, respectively.</p>
Full article ">
3 pages, 212 KiB  
Editorial
Special Issue “Advanced Nanomaterials for Bioimaging”
by Gang Ho Lee
Nanomaterials 2022, 12(14), 2496; https://doi.org/10.3390/nano12142496 - 20 Jul 2022
Cited by 1 | Viewed by 1250
Abstract
Bioimaging currently plays a critical role in medical diagnosis [...] Full article
(This article belongs to the Special Issue Advanced Nanomaterials for Bioimaging)
25 pages, 6988 KiB  
Review
Recent Progress in Flexible Pressure Sensor Arrays
by Yanhao Duan, Shixue He, Jian Wu, Benlong Su and Youshan Wang
Nanomaterials 2022, 12(14), 2495; https://doi.org/10.3390/nano12142495 - 20 Jul 2022
Cited by 38 | Viewed by 7368
Abstract
Flexible pressure sensors that can maintain their pressure sensing ability with arbitrary deformation play an essential role in a wide range of applications, such as aerospace, prosthetics, robotics, healthcare, human–machine interfaces, and electronic skin. Flexible pressure sensors with diverse conversion principles and structural [...] Read more.
Flexible pressure sensors that can maintain their pressure sensing ability with arbitrary deformation play an essential role in a wide range of applications, such as aerospace, prosthetics, robotics, healthcare, human–machine interfaces, and electronic skin. Flexible pressure sensors with diverse conversion principles and structural designs have been extensively studied. At present, with the development of 5G and the Internet of Things, there is a huge demand for flexible pressure sensor arrays with high resolution and sensitivity. Herein, we present a brief description of the present flexible pressure sensor arrays with different transduction mechanisms from design to fabrication. Next, we discuss the latest progress of flexible pressure sensor arrays for applications in human–machine interfaces, healthcare, and aerospace. These arrays can monitor the spatial pressure and map the trajectory with high resolution and rapid response beyond human perception. Finally, the outlook of the future and the existing problems of pressure sensor arrays are presented. Full article
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) A PDMS/MWCNT-based tactile sensor array with coplanar electrodes for crosstalk suppression. Reproduced with permission from Ref. [<a href="#B60-nanomaterials-12-02495" class="html-bibr">60</a>]. Copyright 2016, Luxian Wang et al. (<b>b</b>) A tactile-direction-sensitive and stretchable electronic skin based on human-skin-inspired interlocked microstructures. Reproduced with permission from Ref. [<a href="#B61-nanomaterials-12-02495" class="html-bibr">61</a>]. Copyright 2014, American Chemical Society. (<b>c</b>) A sensor array based on a highly sensitive, broad-range, hierarchically structured reduced graphene oxide/PolyHIPE foam. Reproduced with permission from Ref. [<a href="#B63-nanomaterials-12-02495" class="html-bibr">63</a>]. Copyright 2019, American Chemical Society. (<b>d</b>) A flexible and self-powered dual-parameter temperature–pressure sensor array using microstructure-frame-supported organic thermoelectric materials. Reproduced with permission from Ref. [<a href="#B64-nanomaterials-12-02495" class="html-bibr">64</a>]. Copyright 2015, Fengjiao Zhang et al. (<b>e</b>) A quantum-effect-based flexible and transparent pressure sensor array with ultrahigh sensitivity and sensing density. Reproduced with permission from Ref. [<a href="#B65-nanomaterials-12-02495" class="html-bibr">65</a>]. Copyright 2020, Lan Shi et al. (<b>f</b>) A behavior-learned cross-reactive sensor matrix for intelligent skin perception. Reproduced with permission from Ref. [<a href="#B66-nanomaterials-12-02495" class="html-bibr">66</a>]. Copyright 2020, WILEY-VCH.</p>
Full article ">Figure 2
<p>(<b>a</b>) A tunable, ultrasensitive, and flexible pressure sensor array based on wrinkled microstructures. Reproduced with permission from Ref. [<a href="#B71-nanomaterials-12-02495" class="html-bibr">71</a>]. Copyright 2019, American Chemical Society. (<b>b</b>) A flexible, stretchable, and wearable multifunctional sensor array for static and dynamic strain mapping. Reproduced with permission from Ref. [<a href="#B72-nanomaterials-12-02495" class="html-bibr">72</a>]. Copyright 2015, WILEY-VCH. (<b>c</b>) A skin-inspired highly stretchable and conformable matrix network for multifunctional sensing. Reproduced with permission from Ref. [<a href="#B73-nanomaterials-12-02495" class="html-bibr">73</a>]. Copyright 2018, Qilin Hua et al. (<b>d</b>) A conductive fiber-based ultrasensitive textile pressure sensor. Reproduced with permission from Ref. [<a href="#B74-nanomaterials-12-02495" class="html-bibr">74</a>]. Copyright 2015, WILEY-VCH. (<b>e</b>) A flexible, transparent, sensitive, and crosstalk-free capacitive tactile sensor array based on graphene electrodes and air as a dielectric. Reproduced with permission from Ref. [<a href="#B75-nanomaterials-12-02495" class="html-bibr">75</a>]. Copyright 2017, WILEY-VCH. (<b>f</b>) A graded intrafillable-architecture-based iontronic pressure sensor array with ultra-broad-range high sensitivity. Reproduced with permission from Ref. [<a href="#B76-nanomaterials-12-02495" class="html-bibr">76</a>]. Copyright 2020, Ningning Bai et al.</p>
Full article ">Figure 3
<p>(<b>a</b>) A bionic single-electrode electronic skin unit based on a piezoelectric nanogenerator. Reproduced with permission from Ref. [<a href="#B77-nanomaterials-12-02495" class="html-bibr">77</a>]. Copyright 2018, American Chemical Society. (<b>b</b>) A screen printing of a flexible piezoelectric-based device on polyethylene terephthalate and paper for touch- and force-sensing applications. Reproduced with permission from Ref. [<a href="#B79-nanomaterials-12-02495" class="html-bibr">79</a>]. Copyright 2017, Elsevier. (<b>c</b>) A scalable imprinting of a flexible multiplexed sensor array with distributed piezoelectricity-enhanced micropillars for dynamic tactile sensing. Reproduced with permission from Ref. [<a href="#B80-nanomaterials-12-02495" class="html-bibr">80</a>]. Copyright 2020, WILEY-VCH. (<b>d</b>) A flexible sensor array used for touch applications and mapping of force distribution. Reproduced with permission from Ref. [<a href="#B80-nanomaterials-12-02495" class="html-bibr">80</a>]. Copyright 2020, WILEY-VCH. (<b>e</b>) A skin-inspired piezoelectric tactile sensor array with crosstalk-free row + column electrodes for spatiotemporally distinguishing diverse stimuli. Reproduced with permission from Ref. [<a href="#B81-nanomaterials-12-02495" class="html-bibr">81</a>]. Copyright 2021, Weikang Lin et al.</p>
Full article ">Figure 4
<p>(<b>a</b>) High-resolution electroluminescent imaging of pressure distribution using a piezoelectric nanowire LED array. Reproduced with permission from Ref. [<a href="#B82-nanomaterials-12-02495" class="html-bibr">82</a>]. Copyright 2013, Nature Publishing Group. (<b>b</b>) A flexible and controllable piezo-phototronic pressure-mapping sensor matrix with a ZnO NW/p-Polymer LED array. Reproduced with permission from Ref. [<a href="#B83-nanomaterials-12-02495" class="html-bibr">83</a>]. Copyright 2015, WILEY-VCH. (<b>c</b>) Achieving high-resolution pressure mapping via a flexible GaN/ZnO nanowire LED array with the piezo-phototronic effect. Reproduced with permission from Ref. [<a href="#B84-nanomaterials-12-02495" class="html-bibr">84</a>]. Copyright 2019, Elsevier. (<b>d</b>) A CdS nanorod/organic hybrid LED array and the piezo-phototronic effect of the device for pressure mapping. Reproduced with permission from Ref. [<a href="#B85-nanomaterials-12-02495" class="html-bibr">85</a>]. Copyright 2016, Royal Society of Chemistry. (<b>e</b>) A dynamic pressure mapping of personalized handwriting by a flexible sensor array based on the mechanoluminescence process. Reproduced with permission from Ref. [<a href="#B86-nanomaterials-12-02495" class="html-bibr">86</a>]. Copyright 2015, WILEY-VCH.</p>
Full article ">Figure 5
<p>(<b>a</b>) A triboelectric sensor array for self-powered static and dynamic pressure detection and tactile imaging. Reproduced with permission from Ref. [<a href="#B88-nanomaterials-12-02495" class="html-bibr">88</a>]. Copyright 2013, American Chemical Society. (<b>b</b>) An integrated flexible, waterproof, transparent, and self-powered tactile sensing panel. Reproduced with permission from Ref. [<a href="#B93-nanomaterials-12-02495" class="html-bibr">93</a>]. Copyright 2016, American Chemical Society. (<b>c</b>) A large-area integrated triboelectric sensor array for wireless static and dynamic pressure detection and mapping. Reproduced with permission from Ref. [<a href="#B94-nanomaterials-12-02495" class="html-bibr">94</a>]. Copyright 2019, John Wiley and Sons.</p>
Full article ">Figure 6
<p>(<b>a</b>) A self-powered, high-resolution, and pressure-sensitive triboelectric sensor array for real-time tactile mapping. Reproduced with permission from Ref. [<a href="#B91-nanomaterials-12-02495" class="html-bibr">91</a>]. Copyright 2016, WILEY-VCH. (<b>b</b>) A metal-electrode-free, fully integrated, soft triboelectric sensor array for self-powered tactile sensing. Reproduced with permission from Ref. [<a href="#B92-nanomaterials-12-02495" class="html-bibr">92</a>]. Copyright 2020, Lingyun Wang et al. (<b>c</b>) Electrode structural design of the cross type. Reproduced with permission from Ref. [<a href="#B91-nanomaterials-12-02495" class="html-bibr">91</a>]. Copyright 2016, WILEY-VCH. (<b>d</b>) A highly stretchable, transparent, and self-powered triboelectric tactile sensor with metallized nanofibers. Reproduced with permission from Ref. [<a href="#B95-nanomaterials-12-02495" class="html-bibr">95</a>]. Copyright 2017, WILEY-VCH. (<b>e</b>) A self-powered sensor array for high-resolution pressure sensing. Reproduced with permission from Ref. [<a href="#B97-nanomaterials-12-02495" class="html-bibr">97</a>]. (<b>f</b>) Triboelectrification-enabled touch sensing for self-powered position mapping and dynamic tracking with a flexible and area-scalable sensor array. Reproduced with permission from Ref. [<a href="#B98-nanomaterials-12-02495" class="html-bibr">98</a>]. Copyright 2017, Elsevier.</p>
Full article ">Figure 7
<p>(<b>a</b>) A full-dynamic-range pressure sensor matrix based on dual-mode optical and electrical sensing. Reproduced with permission from Ref. [<a href="#B99-nanomaterials-12-02495" class="html-bibr">99</a>]. Copyright 2017, WILEY-VCH. (<b>b</b>) Diagram of pressure regimes and the relevant applications in our daily lives. Reproduced with permission from Ref. [<a href="#B99-nanomaterials-12-02495" class="html-bibr">99</a>]. Copyright 2017, WILEY-VCH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Electrooculography and tactile perception in a collaborative interface for 3D human–machine interaction. Reproduced with permission from Ref. [<a href="#B110-nanomaterials-12-02495" class="html-bibr">110</a>]. Copyright 2022, American Chemical Society. (<b>b</b>) A scalable tactile glove consisting of a sensor array with 548 elements covering the entire hand. Reproduced with permission from Ref. [<a href="#B67-nanomaterials-12-02495" class="html-bibr">67</a>]. Copyright 2019, Subramanian Sundaram et al. (<b>c</b>) Spatial pressure distribution capability test of a 5 × 5 sensor array using plastic boards that are shaped like the letters “U”, “J”, and “N”. Reproduced with permission from Ref. [<a href="#B111-nanomaterials-12-02495" class="html-bibr">111</a>]. Copyright 2019, WILEY-VCH. (<b>d</b>) An image of a numeric keypad with braille. Reproduced with permission from Ref. [<a href="#B111-nanomaterials-12-02495" class="html-bibr">111</a>]. Copyright 2019, WILEY-VCH.</p>
Full article ">Figure 9
<p>(<b>a</b>) Real-time pressure mapping in a smart insole system based on a sensor array. Reproduced with permission from Ref. [<a href="#B115-nanomaterials-12-02495" class="html-bibr">115</a>]. Copyright 2020, Juan Tao et al. (<b>b</b>) Artificial intelligence toilet (AI-Toilet) for an integrated health monitoring system using smart triboelectric pressure sensor arrays. Reproduced with permission from Ref. [<a href="#B116-nanomaterials-12-02495" class="html-bibr">116</a>]. Copyright 2021, Elsevier. (<b>c</b>) Intrinsically stretchable electronics with ultrahigh deformability to monitor dynamically moving organs. Reproduced with permission from Ref. [<a href="#B119-nanomaterials-12-02495" class="html-bibr">119</a>]. Copyright 2022, Shaolei Wang et al.</p>
Full article ">Figure 10
<p>Conformable, programmable, and step-linear sensor array for large-range wind pressure measurements on a curved surface. Reproduced with permission from Ref. [<a href="#B122-nanomaterials-12-02495" class="html-bibr">122</a>]. Copyright 2020, Science China Press and Springer-Verlag GmbH Germany, part of Springer Nature.</p>
Full article ">
14 pages, 3456 KiB  
Article
Tunable Spin and Orbital Edelstein Effect at (111) LaAlO3/SrTiO3 Interface
by Mattia Trama, Vittorio Cataudella, Carmine Antonio Perroni, Francesco Romeo and Roberta Citro
Nanomaterials 2022, 12(14), 2494; https://doi.org/10.3390/nano12142494 - 20 Jul 2022
Cited by 8 | Viewed by 2046
Abstract
Converting charge current into spin current is one of the main mechanisms exploited in spintronics. One prominent example is the Edelstein effect, namely, the generation of a magnetization in response to an external electric field, which can be realized in systems with lack [...] Read more.
Converting charge current into spin current is one of the main mechanisms exploited in spintronics. One prominent example is the Edelstein effect, namely, the generation of a magnetization in response to an external electric field, which can be realized in systems with lack of inversion symmetry. If a system has electrons with an orbital angular momentum character, an orbital magnetization can be generated by the applied electric field, giving rise to the so-called orbital Edelstein effect. Oxide heterostructures are the ideal platform for these effects due to the strong spin–orbit coupling and the lack of inversion symmetries. Beyond a gate-tunable spin Edelstein effect, we predict an orbital Edelstein effect an order of magnitude larger then the spin one at the (111) LaAlO3/SrTiO3 interface for very low and high fillings. We model the material as a bilayer of t2g orbitals using a tight-binding approach, whereas transport properties are obtained in the Boltzmann approach. We give an effective model at low filling, which explains the non-trivial behaviour of the Edelstein response, showing that the hybridization between the electronic bands crucially impacts the Edelstein susceptibility. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Ti atoms in STO lattice, whose lattice constant is <math display="inline"><semantics> <mrow> <msub> <mi>a</mi> <mn>0</mn> </msub> <mo>=</mo> <mn>0.3905</mn> </mrow> </semantics></math> nm. The blue and yellow dots represent atoms belonging to two non-equivalent planes. (<b>b</b>) Projection of the two non-equivalent planes of Ti over the (111) plane with our choice of primitive vectors <math display="inline"><semantics> <msub> <mover accent="true"> <mi>R</mi> <mo stretchy="false">→</mo> </mover> <mn>1</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mover accent="true"> <mi>R</mi> <mo stretchy="false">→</mo> </mover> <mn>2</mn> </msub> </semantics></math> and <math display="inline"><semantics> <mrow> <mover accent="true"> <mi>a</mi> <mo>˜</mo> </mover> <mo>=</mo> <msqrt> <mrow> <mn>2</mn> <mo>/</mo> <mn>3</mn> </mrow> </msqrt> <msub> <mi>a</mi> <mn>0</mn> </msub> </mrow> </semantics></math>. (<b>c</b>) Band structure along two different directions in the Brillouin zone. The purple benchmark line corresponds to a Lifshitz transition (see <a href="#app3-nanomaterials-12-02494" class="html-app">Appendix C</a>).</p>
Full article ">Figure 2
<p>In-plane spin (<b>upper panel</b>) and orbital angular momentum (<b>lower panel</b>) textures for the three doublets with the chemical potential fixed to the value corresponding to the benchmark line 3 in <a href="#nanomaterials-12-02494-f001" class="html-fig">Figure 1</a>. The red and green arrows represent the mean value of the in-plane component of the operator for the external band, and the blue and purple refer to the internal component. The mean value of the generic operator <span class="html-italic">O</span> is evaluated as <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mi>O</mi> <mo>〉</mo> </mrow> <mo>=</mo> <msqrt> <mrow> <msup> <mrow> <mo>〈</mo> <msub> <mi>O</mi> <mrow> <mover> <mn>1</mn> <mo>¯</mo> </mover> <mn>10</mn> </mrow> </msub> <mo>〉</mo> </mrow> <mn>2</mn> </msup> <mo>+</mo> <msup> <mrow> <mo>〈</mo> <msub> <mi>O</mi> <mrow> <mover> <mn>11</mn> <mo>¯</mo> </mover> <mn>2</mn> </mrow> </msub> <mo>〉</mo> </mrow> <mn>2</mn> </msup> </mrow> </msqrt> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Spin (<b>a</b>) and orbital (<b>b</b>) Edelstein coefficient as a function of the chemical potential. The different colours correspond to the contribution of a specific Kramers doublet.</p>
Full article ">Figure 4
<p>Spin (<b>upper panel</b>) and orbital (<b>lower panel</b>) Edelstein susceptibility projected over the <span class="html-italic">L</span>, <span class="html-italic">M</span>, and <span class="html-italic">U</span> states. The chemical potential <math display="inline"><semantics> <mi>μ</mi> </semantics></math> is fixed at values 1, 2, and 3, referring to <a href="#nanomaterials-12-02494-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure A1
<p>In-plane spin (<b>upper panel</b>) and orbital angular momentum (<b>lower panel</b>) textures for the first and the second doublet with the chemical potential fixed to the value corresponding to the benchmark lines 1 and 2 in <a href="#nanomaterials-12-02494-f001" class="html-fig">Figure 1</a>. The red and green arrows represent the mean value of the in-plane component of the operator for the external band, and the blue and pink refer to the internal component. The mean value of the generic operator <span class="html-italic">O</span> is evaluated as <math display="inline"><semantics> <mrow> <mrow> <mo>〈</mo> <mi>O</mi> <mo>〉</mo> </mrow> <mo>=</mo> <msqrt> <mrow> <msup> <mrow> <mo>〈</mo> <msub> <mi>O</mi> <mrow> <mover> <mn>1</mn> <mo>¯</mo> </mover> <mn>10</mn> </mrow> </msub> <mo>〉</mo> </mrow> <mn>2</mn> </msup> <mo>+</mo> <msup> <mrow> <mo>〈</mo> <msub> <mi>O</mi> <mrow> <mover> <mn>11</mn> <mo>¯</mo> </mover> <mn>2</mn> </mrow> </msub> <mo>〉</mo> </mrow> <mn>2</mn> </msup> </mrow> </msqrt> </mrow> </semantics></math>.</p>
Full article ">Figure A2
<p>Graph of <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>/</mo> <mi>τ</mi> </mrow> </semantics></math> as a function of the chemical potential for a benchmark direction in the Brillouin zone. The dashed line corresponds to the inverse of the scattering time used in the main text. The purple vertical line corresponds to the energy at which a Lifshitz transition occurs for first band (see <a href="#nanomaterials-12-02494-f0A3" class="html-fig">Figure A3</a>).</p>
Full article ">Figure A3
<p>Spin (<b>a</b>) and orbital (<b>b</b>) Edelstein coefficient as a function of the chemical potential using a scattering time model with a point-like impurity. The different colours correspond to the contribution of a specific Kramers doublet, whereas the dashed line corresponds to the total Edelstein response for a constant <math display="inline"><semantics> <mi>τ</mi> </semantics></math>. The inset in panel (<b>a</b>) corresponds to the detail of the Fermi energy contour for the purple chemical potential line.</p>
Full article ">
16 pages, 10763 KiB  
Article
Formation of Nano- and Micro-Scale Surface Features Induced by Long-Range Femtosecond Filament Laser Ablation
by Joerg Schille, Jose R. Chirinos, Xianglei Mao, Lutz Schneider, Matthias Horn, Udo Loeschner and Vassilia Zorba
Nanomaterials 2022, 12(14), 2493; https://doi.org/10.3390/nano12142493 - 20 Jul 2022
Cited by 2 | Viewed by 2223
Abstract
In this work, we study the characteristics of femtosecond-filament-laser–matter interactions and laser-induced periodic surface structures (LIPSS) at a beam-propagation distance up to 55 m. The quantification of the periodicity of filament-induced self-organized surface structures was accomplished by SEM and AFM measurements combined with [...] Read more.
In this work, we study the characteristics of femtosecond-filament-laser–matter interactions and laser-induced periodic surface structures (LIPSS) at a beam-propagation distance up to 55 m. The quantification of the periodicity of filament-induced self-organized surface structures was accomplished by SEM and AFM measurements combined with the use of discrete two-dimensional fast Fourier transform (2D-FFT) analysis, at different filament propagation distances. The results show that the size of the nano-scale surface features increased with ongoing laser filament processing and, further, periodic ripples started to form in the ablation-spot center after irradiation with five spatially overlapping pulses. The effective number of irradiating filament pulses per spot area affected the developing surface texture, with the period of the low spatial frequency LIPSS reducing notably at a high pulse number. The high regularity of the filament-induced ripples was verified by the demonstration of the angle-of-incidence-dependent diffraction of sunlight. This work underlines the potential of long-range femtosecond filamentation for energy delivery at remote distances, with suppressed diffraction and long depth focus, which can be used in biomimetic laser surface engineering and remote-sensing applications. Full article
Show Figures

Figure 1

Figure 1
<p>Experimental setup to launch the filament at long distances. Filament-induced LIPSS were studied across the range of 50–55 m from the laser exit. The laser beam was folded a few times before the filament was formed using gold-coated mirrors to reach the 55-meter target position inside the laboratory.</p>
Full article ">Figure 2
<p>Optical microscopy images showing the femtosecond-filament beam-pointing stability on the substrate surface when the substrate was 55 m from the laser source (approximately 5-meter filament distance), as a function of number of pulses. The number of irradiated laser-filament pulses in this case varied between 1 and 100 (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 3
<p>Increase in the filament-ablation-spot diameter as a function of pulse number and substrate position in the filament, which resulted from beam fluctuation in the propagating filament. The insert shows the ablation-spot diameters for 1- and 2-filament pulses in a magnified view.</p>
Full article ">Figure 4
<p>SEM micrographs and 2D-FFT analysis showing the development of the surface texture in the center of the laser-filament-made ablation spot at 5-meter substrate position in the filament path (55-meter beam propagation). The number of filament pulses varied.</p>
Full article ">Figure 5
<p>AFM analysis of un-textured (<b>left</b>) vs. laser-filament-textured steel surfaces achieved by irradiating 1 (<b>center</b>) and 20 (<b>right</b>) filament pulses at 5-m substrate position in the filament path (55-m beam propagation). The positions of the cross-section profiles shown in this figure’s lower part are indicated by red lines in the 2D height maps presented in upper part of the figure.</p>
Full article ">Figure 6
<p>Overview of the surface textures developing in the center of the laser-filament-induced ablation spots at 2–5-m substrate positions in the filament propagation path and 52–55-m corresponding beam-propagation distances. The pulse number and substrate position in the filament varied.</p>
Full article ">Figure 7
<p>SEM micrographs showing the surface textures developing at different positions within the laser-filament-induced ablation spots (5 m (<b>top</b>, <b>center</b>) and 4 m (<b>bottom</b>) in filament path, or 55 and 54 m of total beam-propagation distance). The number of irradiated filament pulses and the position of the substrate surface in the filament varied.</p>
Full article ">Figure 8
<p>SEM micrographs (<b>upper</b> row), 2D-FFT map (<b>middle</b> row), and wave number profile along the <span class="html-italic">k</span>-axis (<b>bottom</b> row), as obtained in the outer areas of laser-filament-induced ablation spots at 4-m substrate position (54-m beam propagation).</p>
Full article ">Figure 9
<p>SEM micrographs (<b>upper</b> row), 2D-FFT maps (<b>middle</b> row), and wave-number profiles along the <span class="html-italic">k</span>-axis (<b>bottom</b> row), as obtained in the outer areas of laser-filament-made ablation spots at 5-m substrate position (55-m beam propagation).</p>
Full article ">Figure 10
<p>(<b>Left</b>): Ripple period and DLOA determined in the outer areas of laser-filament ablation spots produced at 4-m and 5-m substrate positions in the filament (54 m and 55 m of beam propagation); the number of irradiated filament pulses varied. (<b>Right</b>): Filament-made ripple texture as a diffraction grating for producing rainbow colors from sunlight.</p>
Full article ">
10 pages, 454 KiB  
Article
Band Structure Near the Dirac Point in HgTe Quantum Wells with Critical Thickness
by Alexey Shuvaev, Vlad Dziom, Jan Gospodarič, Elena G. Novik, Alena A. Dobretsova, Nikolay N. Mikhailov, Ze Don Kvon and Andrei Pimenov
Nanomaterials 2022, 12(14), 2492; https://doi.org/10.3390/nano12142492 - 20 Jul 2022
Cited by 7 | Viewed by 1722
Abstract
Mercury telluride (HgTe) thin films with a critical thickness of 6.5 nm are predicted to possess a gapless Dirac-like band structure. We report a comprehensive study on gated and optically doped samples by magnetooptical spectroscopy in the THz range. The quasi-classical analysis of [...] Read more.
Mercury telluride (HgTe) thin films with a critical thickness of 6.5 nm are predicted to possess a gapless Dirac-like band structure. We report a comprehensive study on gated and optically doped samples by magnetooptical spectroscopy in the THz range. The quasi-classical analysis of the cyclotron resonance allowed the mapping of the band dispersion of Dirac charge carriers in a broad range of electron and hole doping. A smooth transition through the charge neutrality point between Dirac holes and electrons was observed. An additional peak coming from a second type of holes with an almost density-independent mass of around 0.04m0 was detected in the hole-doping range and attributed to an asymmetric spin splitting of the Dirac cone. Spectroscopic evidence for disorder-induced band energy fluctuations could not be detected in present cyclotron resonance experiments. Full article
(This article belongs to the Special Issue Semiconductor Quantum Wells and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Magneto-transmission of a circularly polarized light in HgTe quantum wells with critical thickness at <math display="inline"><semantics> <mrow> <mi>T</mi> <mo>=</mo> <mn>1.8</mn> </mrow> </semantics></math> K and <math display="inline"><semantics> <mrow> <mi>ν</mi> <mo>=</mo> <mn>950</mn> </mrow> </semantics></math> GHz. (<b>a</b>) Sample #1 with gate (gate voltage is given at curves). (<b>b</b>) Sample #2 without gate. The charge density here is varied via light illumination with illumination time as indicated. Negative and positive magnetic fields correspond to negatively- and positively-charged quasiparticles, respectively. Acronyms: D.E.—Dirac electrons, D.H.—Dirac holes, M.H.—massive holes (see text for details). The inset in (<b>b</b>) demonstrates schematically the expected Fermi surface for different doping levels.</p>
Full article ">Figure 2
<p>Parameters of the charge carriers as extracted from the Drude fits of the cyclotron resonance. (<b>a</b>–<b>c</b>) Sample #1 with gate as function of gate voltage, (<b>d</b>–<b>f</b>) Sample #2 without gate as function of illumination time. (<b>a</b>,<b>d</b>) Cyclotron mass, (<b>b</b>,<b>e</b>) Charge density, (<b>c</b>,<b>f</b>) Mobility. Acronyms: blue circles-Dirac electrons (D.E.), brown circles-Dirac holes (D.H.), red triangles-massive holes (M.H.).</p>
Full article ">Figure 3
<p>Cyclotron mass <math display="inline"><semantics> <msub> <mi>m</mi> <mi>c</mi> </msub> </semantics></math> as a function of density <span class="html-italic">n</span>. Note the square root scale of the horizontal axis. The data from the gated sample #1 are shown by filled symbols, from the sample #2 without gate-by open symbols. Dirac electrons and holes are given by dark blue and brown circles, respectively. Massive holes are given by red triangles. The blue and brown lines are the predictions by the <math display="inline"><semantics> <mrow> <mi mathvariant="bold">k</mi> <mo>·</mo> <mi mathvariant="bold">p</mi> </mrow> </semantics></math> model for Dirac electrons and holes, respectively. The difference between solid and dashed lines is due to splitting in the asymmetric potential and to the bulk inversion asymmetry.</p>
Full article ">Figure 4
<p>Band structure of HgTe films with critical thickness. Symbols are experimental data from two samples, lines—<math display="inline"><semantics> <mrow> <mi mathvariant="bold">k</mi> <mo>·</mo> <mi mathvariant="bold">p</mi> </mrow> </semantics></math> theory. The symbol notations are the same as in <a href="#nanomaterials-12-02492-f003" class="html-fig">Figure 3</a>. The solid and dashed lines are split bands due to asymmetric potential and to bulk inversion asymmetry. The different slope of electrons and holes is due to a lifted hole degeneracy. The inset shows the overview over the data on a large energy scale. The flattening of the hole dispersion at high values of <span class="html-italic">k</span> is due to approaching of the hole pockets around <math display="inline"><semantics> <mrow> <mi>k</mi> <mo>∼</mo> <mn>0.5</mn> </mrow> </semantics></math> nm<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>, <math display="inline"><semantics> <mrow> <mi>E</mi> <mo>∼</mo> <mo>−</mo> <mn>15</mn> </mrow> </semantics></math> meV.</p>
Full article ">
18 pages, 4488 KiB  
Article
Adsorptive Removal of Naproxen from Water Using Polyhedral Oligomeric Silesquioxane (POSS) Covalent Organic Frameworks (COFs)
by Suleiman Bala, Che Azurahanim Che Abdullah, Mohamed Ibrahim Mohamed Tahir and Mohd Basyaruddin Abdul Rahman
Nanomaterials 2022, 12(14), 2491; https://doi.org/10.3390/nano12142491 - 20 Jul 2022
Cited by 12 | Viewed by 2355
Abstract
Covalent organic frameworks are porous crystalline compounds made up of organic material bonded together by strong reversible covalent bonds (these are novel types of materials which have the processability of extended or repeated structures with high performance, like those of thermosets and thermoplastics [...] Read more.
Covalent organic frameworks are porous crystalline compounds made up of organic material bonded together by strong reversible covalent bonds (these are novel types of materials which have the processability of extended or repeated structures with high performance, like those of thermosets and thermoplastics that produce high surface coverage). These have a long-term effect on an arrangement’s geometry and permeability. These compounds are entirely made up of light elements like H, B, C, N, O and Si. Pharmaceuticals and personal care products (PPCPs) have emerged as a new threatened species. A hazardous substance known as an “emerging toxin,” such as naproxen, is one that has been established or is generated in sufficient amounts in an environment, creating permanent damage to organisms. COF-S7, OAPS and 2-methylanthraquionone(2-MeAQ), and COF-S12, OAPS and terephthalaldehyde (TPA) were effectively synthesized by condensation (solvothermal) via a Schiff base reaction (R1R2C=NR′), with a molar ratio of 1:8 for OAPS to linker (L1 and L2), at a temperature of 125 °C and 100 °C for COF-S7 and COF-S12, respectively. The compounds obtained were assessed using several spectroscopy techniques, which revealed azomethine C=N bonds, aromatic carbon environments via solid 13C and 29Si NMR, the morphological structure and porosity, and the thermostability of these materials. The remedied effluent was investigated, and a substantial execution was noted in the removal ability of the naproxen over synthesized materials, such as 70% and 86% at a contact time of 210 min and 270 min, respectively, at a constant dose of 0.05 g and pH 7. The maximum adsorption abilities of the substances were found to be 35 mg/g and 42 mg/g. The pH result implies that there is stable exclusion with a rise in pH to 9. At pH 9, the drop significance was attained for COF-S7 with the exception of COF-S12, which was detected at pH 11, due to the negative Foster charge, consequent to the repulsion among the synthesized COFs and naproxen solution. From the isotherms acquired (Langmuir and Freundlich), the substances displayed a higher value (close to 1) of correlation coefficient (R2), which showed that the substances fit into the Freundlich isotherm (heterogenous process), and the value of heterogeneity process (n) achieved (less than 1) specifies that the adsorption is a chemical process. Analysis of the as-prepared composites revealed remarkable reusability in the elimination of naproxen by adsorption. Due to its convenience of synthesis, significant adsorption effectiveness, and remarkable reusability, the as-synthesized COFs are expected to be able to be used as potential adsorbents for eliminating AIDs from water. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>PXRD spectra of as-synthesized and simulated (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12.</p>
Full article ">Figure 2
<p>PXRD spectra of as-synthesized COFs with different solvents for chemical stability study (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12.</p>
Full article ">Figure 3
<p>Stack plot comparing, (<b>a</b>). (black) ONPS and (red) OAPS (<b>b</b>) the FTIR spectra of as-synthesized (black curve) COF-S7, (red curve) OAPS and (blue curve) 2-MeAQ (<b>c</b>) (black curve) COF-S12, (red curve) OAPS, and (blue curve) TPA.</p>
Full article ">Figure 4
<p><sup>1</sup>H-NMR spectra of (<b>a</b>) ONPS and (<b>b</b>) OAPS.</p>
Full article ">Figure 5
<p>(<b>a</b>) CP-MAS <sup>13</sup>C solid-state NMR spectra of OAPS (<b>b</b>) CP-MAS <sup>29</sup>Si solid-state NMR spectra of OAPS.</p>
Full article ">Figure 6
<p>CP-MAS <sup>13</sup>C solid-state NMR spectrum of (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12, and CP-MAS <sup>29</sup>Si solid-state spectrum of (<b>c</b>) COF-S7 and (<b>d</b>) COF-S12.</p>
Full article ">Figure 7
<p>TGA thermogram of (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12 at heating rate of 10 °C min<sup>−1</sup>.</p>
Full article ">Figure 8
<p>FESEM images of COF-S7 (<b>a</b>,<b>b</b>), at 100× <span class="html-italic">g</span> and 50× <span class="html-italic">g</span> magnifications, COF-S12 (<b>c</b>,<b>d</b>) at 25× <span class="html-italic">g</span> and 50× <span class="html-italic">g</span> magnifications under 5.0 kV accelerating voltage.</p>
Full article ">Figure 9
<p>EDX spectra and weight percentages of identified elements (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12.</p>
Full article ">Figure 10
<p>N<sub>2</sub> isotherm of (<b>a</b>) COF-S7 and (<b>b</b>) COF-S12 at 77 K.</p>
Full article ">Figure 11
<p>(<b>a</b>) NAP removal efficiency of COF-S7 and COF-S12 (<b>b</b>) NAP adsorption capacity of COF-S7 and COFS12 (<b>c</b>) Effect of pH on the adsorbed amounts of NAP over synthesized COFs at fixed dosage (0.05 g) and constant time, (270 min).</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>) Langmuir and Freundlich isotherm plot COF-S7 and (<b>c</b>,<b>d</b>) Langmuir and Freundlich isotherm plot of COF-S12, respectively, in naproxen removal at pH 7, dosage 0.05 g.</p>
Full article ">Figure 13
<p>Adsorbed amount of naproxen over adsorbents (COF-S7 and COF-S12).</p>
Full article ">Scheme 1
<p>The kinetics of naproxen removal in COFs.</p>
Full article ">Scheme 2
<p>Nitration of (<b>a</b>) yields (<b>b</b>), and reduction to produce (<b>c</b>). Condensation of (<b>c</b>,<b>d</b>) and (<b>c</b>,<b>e</b>) lead to the production of (molecular units) which will join together the tetrahedral building blocks D and E to produce the diamond like structures of POSS COF-S7 (<b>f</b>) and COF-S12, (<b>g</b>) single framework (space filling, C brown, O red, N blue, and Si yellow). Hydrogens were omitted for clarity.</p>
Full article ">
21 pages, 3573 KiB  
Article
Preliminary Evaluation of Iron Oxide Nanoparticles Radiolabeled with 68Ga and 177Lu as Potential Theranostic Agents
by Evangelia-Alexandra Salvanou, Argiris Kolokithas-Ntoukas, Christos Liolios, Stavros Xanthopoulos, Maria Paravatou-Petsotas, Charalampos Tsoukalas, Konstantinos Avgoustakis and Penelope Bouziotis
Nanomaterials 2022, 12(14), 2490; https://doi.org/10.3390/nano12142490 - 20 Jul 2022
Cited by 12 | Viewed by 2310
Abstract
Theranostic radioisotope pairs such as Gallium-68 (68Ga) for Positron Emission Tomography (PET) and Lutetium-177 (177Lu) for radioisotopic therapy, in conjunction with nanoparticles (NPs), are an emerging field in the treatment of cancer. The present work aims to demonstrate the [...] Read more.
Theranostic radioisotope pairs such as Gallium-68 (68Ga) for Positron Emission Tomography (PET) and Lutetium-177 (177Lu) for radioisotopic therapy, in conjunction with nanoparticles (NPs), are an emerging field in the treatment of cancer. The present work aims to demonstrate the ability of condensed colloidal nanocrystal clusters (co-CNCs) comprised of iron oxide nanoparticles, coated with alginic acid (MA) and stabilized by a layer of polyethylene glycol (MAPEG) to be directly radiolabeled with 68Ga and its therapeutic analog 177Lu. 68Ga/177Lu- MA and MAPEG were investigated for their in vitro stability. The biocompatibility of the non-radiolabeled nanoparticles, as well as the cytotoxicity of MA, MAPEG, and [177Lu]Lu-MAPEG were assessed on 4T1 cells. Finally, the ex vivo biodistribution of the 68Ga-labeled NPs as well as [177Lu]Lu-MAPEG was investigated in normal mice. Radiolabeling with both radioisotopes took place via a simple and direct labelling method without further purification. Hemocompatibility was verified for both NPs, while MTT studies demonstrated the non-cytotoxic profile of the nanocarriers and the dose-dependent toxicity for [177Lu]Lu-MAPEG. The radiolabeled nanoparticles mainly accumulated in RES organs. Based on our preliminary results, we conclude that MAPEG could be further investigated as a theranostic agent for PET diagnosis and therapy of cancer. Full article
(This article belongs to the Special Issue Nanoparticles in Diagnostic and Therapeutic Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Hydrodynamic diameters (<span class="html-italic">D</span><sub>h</sub>’s) and (<b>b</b>) ζ-potentials of the plain (MA) and PEGylated (MAPEG) co-CNCs MIONs.</p>
Full article ">Figure 2
<p>Representative radio-TLC chromatograph of [<sup>68</sup>Ga]Ga-MA or [<sup>68</sup>Ga]Ga-MAPEG obtained after 30 min of incubation at 75 °C.</p>
Full article ">Figure 3
<p>Size distribution of MIONs before (MAPEG) and after (<sup>68</sup>Ga-MAPEG) radiolabeling with <sup>68</sup>Ga.</p>
Full article ">Figure 4
<p>Radiochemical stability of [<sup>68</sup>Ga]Ga-MA and [<sup>68</sup>Ga]Ga-MAPEG at RT and in human serum (x axis not in scale).</p>
Full article ">Figure 5
<p>Representative radio-TLC chromatograph of [<sup>177</sup>Lu]Lu-MA or [<sup>177</sup>Lu]Lu-MAPEG after 30 min of incubation at 75 °C.</p>
Full article ">Figure 6
<p>Size distribution of MIONs before (MAPEG) and after (<sup>177</sup>Lu-MAPEG) radiolabeling with <sup>177</sup>Lu.</p>
Full article ">Figure 7
<p>Radiochemical stability of [<sup>177</sup>Lu]Lu-MA and [<sup>177</sup>Lu]Lu-MAPEG at RT and in human serum (x axis not in scale).</p>
Full article ">Figure 8
<p>Hemolytic effect of different concentrations of MA and MAPEG. Positive control: 500 μL H<sub>2</sub>O + 15 μL of RBCs; negative control: 500 μL PBS + 15 μL of RBCs.</p>
Full article ">Figure 9
<p>MTT assay of MA against the 4T1 cell line after 24, 48, and 72 h. Mean values (<span class="html-italic">n</span> = 3) and the SD (bars) are shown (x axis not in scale).</p>
Full article ">Figure 10
<p>MTT assay of MAPEG against the 4T1 cell line after 24, 48 and 72 h. Mean values (<span class="html-italic">n</span> = 3) and the SD (bars) are shown (x axis not in scale).</p>
Full article ">Figure 11
<p>MTT assay of [<sup>177</sup>Lu]Lu-MAPEG against the 4T1 cell line after 24 h. Mean values (<span class="html-italic">n</span> = 3) and the SD (bars) are shown (x axis not in scale).</p>
Full article ">Figure 12
<p>Biodistribution of [<sup>68</sup>Ga]Ga-MA in normal mice expressed as % IA/g (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 13
<p>Biodistribution of [<sup>68</sup>Ga]Ga-MAPEG in normal mice expressed as % IA/g (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 14
<p>Biodistribution of [<sup>177</sup>Lu]Lu-MAPEG in normal mice expressed as % IA/g (<span class="html-italic">n</span> = 3).</p>
Full article ">
14 pages, 6542 KiB  
Article
Mechanical Load-Induced Atomic-Scale Deformation Evolution and Mechanism of SiC Polytypes Using Molecular Dynamics Simulation
by Haoxiang Wang, Shang Gao, Renke Kang, Xiaoguang Guo and Honggang Li
Nanomaterials 2022, 12(14), 2489; https://doi.org/10.3390/nano12142489 - 20 Jul 2022
Cited by 22 | Viewed by 2548
Abstract
Silicon carbide (SiC) is a promising semiconductor material for making high-performance power electronics with higher withstand voltage and lower loss. The development of cost-effective machining technology for fabricating SiC wafers requires a complete understanding of the deformation and removal mechanism. In this study, [...] Read more.
Silicon carbide (SiC) is a promising semiconductor material for making high-performance power electronics with higher withstand voltage and lower loss. The development of cost-effective machining technology for fabricating SiC wafers requires a complete understanding of the deformation and removal mechanism. In this study, molecular dynamics (MD) simulations were carried out to investigate the origins of the differences in elastic–plastic deformation characteristics of the SiC polytypes, including 3C-SiC, 4H-SiC and 6H-SiC, during nanoindentation. The atomic structures, pair correlation function and dislocation distribution during nanoindentation were extracted and analyzed. The main factors that cause elastic–plastic deformation have been revealed. The simulation results show that the deformation mechanisms of SiC polytypes are all dominated by amorphous phase transformation and dislocation behaviors. Most of the amorphous atoms recovered after completed unload. Dislocation analysis shows that the dislocations of 3C-SiC are mainly perfect dislocations during loading, while the perfect dislocations in 4H-SiC and 6H-SiC are relatively few. In addition, 4H-SiC also formed two types of stacking faults. Full article
(This article belongs to the Special Issue Micro/Nano-Machining: Fundamentals and Recent Advances)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) MD simulation model of nanoindentation for three types of SiC single crystal: (<b>b</b>) 3C-SiC, (<b>c</b>) 4H-SiC, and (<b>d</b>) 6H-SiC, respectively.</p>
Full article ">Figure 2
<p>Motion parameter of diamond indenter.</p>
Full article ">Figure 3
<p>Cross-sectional view of nanoindentation for three types of SiC single crystal at completed load and completed unload. (<b>a</b>) (100) plane in 3C-SiC, (<b>b</b>) (1-210) plane in 4H-SiC, and (<b>c</b>) (1-210) plane in 6H-SiC, respectively.</p>
Full article ">Figure 4
<p>Pair correlation analysis for three types of SiC single crystal. (<b>a</b>) 3C-SiC, (<b>b</b>) 4H-SiC, and (<b>c</b>) 6H-SiC, respectively.</p>
Full article ">Figure 5
<p>The number of amorphous atoms for three types of SiC single crystal at completed load and completed unload.</p>
Full article ">Figure 6
<p>Coordination analysis for three types of SiC single crystal at completed load and completed unload. (<b>a</b>) 3C-SiC, (<b>b</b>) 4H-SiC, and (<b>c</b>) 6H-SiC, respectively.</p>
Full article ">Figure 7
<p>Dislocation line length of three types of SiC single crystal during loading and unloading process: (<b>a</b>) 3C-SiC, (<b>b</b>) 4H-SiC, and (<b>c</b>) 6H-SiC, respectively.</p>
Full article ">Figure 8
<p>Dislocation distribution of 3C-SiC at different indentation depths: (<b>a</b>) <span class="html-italic">h</span> = 1.3 nm, (<b>b</b>) <span class="html-italic">h</span> = 1.6 nm, (<b>c</b>) <span class="html-italic">h</span> = 2.0 nm, and (<b>d</b>) <span class="html-italic">h</span> = 2.6 nm, respectively.</p>
Full article ">Figure 9
<p>Dislocation distribution of 4H-SiC at different indentation depths: (<b>a</b>) <span class="html-italic">h</span> = 1.3 nm, (<b>b</b>) <span class="html-italic">h</span> = 1.8 nm, (<b>c</b>) <span class="html-italic">h</span> = 2.0 nm, and (<b>d</b>) <span class="html-italic">h</span> = 2.5 nm, respectively.</p>
Full article ">Figure 10
<p>Dislocation distribution of 6H-SiC at different indentation depths: (<b>a</b>) <span class="html-italic">h</span> = 1.1 nm, (<b>b</b>) <span class="html-italic">h</span> = 1.6 nm, (<b>c</b>) <span class="html-italic">h</span> = 2.0 nm, and (<b>d</b>) <span class="html-italic">h</span> = 2.5 nm, respectively.</p>
Full article ">Figure 11
<p>Schematic diagram of the stacking faults formation in 4H-SiC. (<b>a</b>) 4h hexagonal crystal structure, (<b>b</b>) 2H hexagonal crystal, and (<b>c</b>) 3C cubic crystal.</p>
Full article ">Figure 12
<p>Temperature of SiC polytypes during indentation.</p>
Full article ">
13 pages, 4211 KiB  
Article
A Durable Magnetic Superhydrophobic Melamine Sponge: For Solving Complex Marine Oil Spills
by Hanmo Si, Qingwang Liu, Zhenzhong Fan, Biao Wang, Qilei Tong and Mengqi Lin
Nanomaterials 2022, 12(14), 2488; https://doi.org/10.3390/nano12142488 - 20 Jul 2022
Cited by 7 | Viewed by 2114
Abstract
The problem of offshore oil leakage has wreaked havoc on the environment and people’s health. A simple and environmentally friendly impregnation method combined with marine mussel bionics was used to address this issue. Using the viscosity of polydopamine (PDA), nano- Fe3O [...] Read more.
The problem of offshore oil leakage has wreaked havoc on the environment and people’s health. A simple and environmentally friendly impregnation method combined with marine mussel bionics was used to address this issue. Using the viscosity of polydopamine (PDA), nano- Fe3O4 and WS2 adhered to the framework of the melamine sponge (MS), and then the magnetic sponge was modified with n-octadecanethiol (OTD), and finally the superhydrophobic magnetic melamine sponge (mMS) was prepared. The modified sponge has superhydrophobicity (WCA, 156.8° ± 1.18°), high adsorbability (40~100 g°g−1), recyclability (oil adsorbability remains essentially unchanged after 25 cycles), efficient oil–water separation performance (>98%), and can quickly separate oil on the water’s surface and underwater. Furthermore, the modified sponge exhibits excellent stability and durability under harsh operating conditions such as strong sunlight, strong acid, strong alkali, and high salt, and can control the direction of the sponge’s movement by loading a magnetic field. To summarize, mMS has many potential applications as a new magnetic adsorption material for dealing with complex offshore oil spill events. Full article
(This article belongs to the Special Issue Current Review in Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of preparation of mMS.</p>
Full article ">Figure 2
<p>SEM images of MS (<b>a</b>,<b>b</b>) and mMS (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 3
<p>EDS diagram of MS (<b>a</b>) and mMS (<b>b</b>); (<b>c</b>) EDS elemental mapping of mMS in the selected area.</p>
Full article ">Figure 4
<p>FTIR diagram of MS and mMS.</p>
Full article ">Figure 5
<p>Images of water contact angle of MS (<b>a</b>) and mMS (<b>b</b>); water droplets of methyl orange stain on the surface of MS and mMS (<b>c</b>);MS and mMS immersed in water (<b>d</b>); silver mirror phenomenon formed by pressing mMS into water under external force (<b>e</b>); water and kerosene drops on the surface of mMS (<b>f</b>).</p>
Full article ">Figure 6
<p>Adsorption ratio of mMS to different oils and organic solvents (<b>a</b>); adsorption ratio of mMS to kerosene, diesel and corn germ oil after multiple pressing cycles (<b>b</b>).</p>
Full article ">Figure 7
<p>Oil–water separation efficiency of mMS in dynamic oil–water mixtures (<b>a</b>); a simple device separates water-CCl<sub>4</sub> mixture (<b>b</b>); adsorption of crude oil on simulated seawater surface by mMS (<b>c</b>).</p>
Full article ">Figure 8
<p>The influence of pH (<b>a</b>), temperature (<b>b</b>), NaCl concentration (<b>c</b>), and strong light (<b>d</b>) on the water contact Angle of mMS.</p>
Full article ">Figure 9
<p>Use a magnet to control the moving direction of mMS to absorb kerosene on the water surface.</p>
Full article ">
9 pages, 3581 KiB  
Article
Comparative Research on the Thermophysical Properties of Nano-Sized La2(Zr0.7Ce0.3)2O7 Synthesized by Different Routes
by Yue Wang, Bohuai Shao, Boyan Fu, Binglin Zou and Chunjie Wang
Nanomaterials 2022, 12(14), 2487; https://doi.org/10.3390/nano12142487 - 20 Jul 2022
Cited by 2 | Viewed by 1531
Abstract
La2(Zr0.7Ce0.3)2O7 has been regarded as an ideal candidate for the next generation of thermal barrier coatings (TBCs) due to its prominent superiority. In this paper, the nano-sized La2(Zr0.7Ce0.3) [...] Read more.
La2(Zr0.7Ce0.3)2O7 has been regarded as an ideal candidate for the next generation of thermal barrier coatings (TBCs) due to its prominent superiority. In this paper, the nano-sized La2(Zr0.7Ce0.3)2O7 was synthesized using two different synthetic routes: sol-gel and hydrothermal processes. Various techniques were utilized to assess the differences in the relevant thermophysical properties created by the different synthetic methods. According to the investigations, both samples exhibited pyrochlore structures with an excellent thermal stability. The sample synthesized via the hydrothermal method showed a more uniform particle size and morphology than that obtained through the sol-gel technique. The former also possessed a better sinter-resistance property, a more outstanding TEC (thermal expansion coefficient) and thermal conductivity, and a larger activation energy for crystal growth than the latter. The micro-strain of both samples showed an interesting change as the temperature increased, and 1200 °C was the turning point. Additionally, relative mechanisms were discussed in detail. Full article
Show Figures

Figure 1

Figure 1
<p>TG−DSC curves of dried powders: (<b>a</b>) LZ7C3−HT; (<b>b</b>) LZ7C3−SG.</p>
Full article ">Figure 2
<p>XRD patterns of LZ7C3−HT (<b>a</b>) and LZ7C3−SG (<b>b</b>) calcined at different temperatures; XRD patterns (<b>c</b>) and Raman spectra (<b>d</b>) of both samples calcined at 1300 °C.</p>
Full article ">Figure 3
<p>Relative densities (<b>a</b>) and volume shrinkages (<b>b</b>) of LZ7C3−HT and LZ7C3−SG bodies under different temperatures.</p>
Full article ">Figure 4
<p>SEM images of as-prepared powders and compacted bodies calcined at 1400 °C of LZ7C3−HT (<b>a</b>,<b>c</b>) and LZ7C3−SG (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>Plots of βcosθ vs. 2sinθ for various peaks (hkl) under different temperatures: (<b>a</b>) LZ7C3−HT; (<b>b</b>) LZ7C3−SG.</p>
Full article ">Figure 6
<p>Relation between the crystal size and temperature plots of ln (<span class="html-italic">D</span><sub>t</sub>/<span class="html-italic">D</span><sub>0</sub>) against 1000/T.</p>
Full article ">Figure 7
<p>TECs (<b>a</b>) and thermal conductivities (<b>b</b>) of LZ7C3−HT and LZ7C3−SG as functions of temperature.</p>
Full article ">
11 pages, 2676 KiB  
Article
Mechanism for the Intercalation of Aniline Cations into the Interlayers of Graphite
by Yifan Guo, Ying Li, Wei Wei, Junhua Su, Jinyang Li, Yanlei Shang, Yong Wang, Xiaoling Xu, David Hui and Zuowan Zhou
Nanomaterials 2022, 12(14), 2486; https://doi.org/10.3390/nano12142486 - 20 Jul 2022
Cited by 4 | Viewed by 2148
Abstract
The dynamic behaviors of aniline cation (ANI+) intercalating into graphite interlayers are systematically studied by experimental studies and multiscale simulations. The in situ intercalation polymerization designed by response surface methods implies the importance of ultrasonication for achieving the intercalation of ANI [...] Read more.
The dynamic behaviors of aniline cation (ANI+) intercalating into graphite interlayers are systematically studied by experimental studies and multiscale simulations. The in situ intercalation polymerization designed by response surface methods implies the importance of ultrasonication for achieving the intercalation of ANI+. Molecular dynamics and quantum chemical simulations prove the adsorption of ANI+ onto graphite surfaces by cation–π electrostatic interactions, weakening the π–π interactions between graphene layers. The ultrasonication that follows breaks the hydrated ANI+ clusters into individual ANI+. Thus, the released positive charges of these dissociative cations and reduced steric hindrance significantly improve their intercalation ability. With the initial kinetic energy provided by ultrasonic field, the activated ANI+ are able to intercalate into the interlayer of graphite. This work demonstrates the intercalation behaviors of ANI+, which provides an opportunity for investigations regarding organic-molecule-intercalated graphite compounds. Full article
(This article belongs to the Special Issue Theoretical Calculation and Molecular Modeling of Nanomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of the exfoliation of graphene (Y) to ultrasonication power (X1) and time (X2) when the graphite is (<b>a</b>) small flaky graphite (X3 = SfG), (<b>b</b>) large flaky graphite (X3 = LfG), and (<b>c</b>) microcrystalline graphite (X3 = MG).</p>
Full article ">Figure 2
<p>MD simulations of ANI<sup>+</sup> adsorbed on the surface of bilayer graphene. (<b>a</b>–<b>c</b>) Snapshots of the adsorption process with simulated times of (<b>a</b>) 0 ns, (<b>b</b>) 5 ns, and (<b>c</b>) 10 ns. ANI<sup>+</sup> within 0.5 nm from graphene is presented using the van der Waals surface model (red), as the others were the transparent line model. H<sub>2</sub>O and Cl<sup>−</sup> are hidden to facilitate observation. (<b>d</b>) Radial distribution function between the centroid of ANI<sup>+</sup> and carbon atoms in bilayer graphene, and (insert) their adsorption model. (<b>e</b>,<b>f</b>) van der Waals interaction energy between (<b>e</b>) ANI<sup>+</sup> and bilayer graphene, and (<b>f</b>) graphene layers in bilayer graphene.</p>
Full article ">Figure 3
<p>First-principles simulations of the adsorption models. Electronic density difference between bilayer graphene and (<b>a</b>,<b>c</b>) ANI<sup>+</sup>, and (<b>b</b>,<b>d</b>) ANIm with an isovalue of 0.003.</p>
Full article ">Figure 4
<p>RESP-mapped van der Waals surface. (<b>a</b>) ANI<sup>+</sup>, (<b>b</b>) ANIm.</p>
Full article ">Figure 5
<p>Interactions between ANI<sup>+</sup> and H<sub>2</sub>O in solution. (<b>a</b>) Solvent-accessible surface area of ANI<sup>+</sup> in solution; (<b>b</b>) number of hydrogen bonds formed between ANI<sup>+</sup> and H<sub>2</sub>O; (<b>c</b>) radial distribution function between ANI<sup>+</sup> and H<sub>2</sub>O; (<b>d</b>) RESP mapped van der Waals surface of ANI<sup>+</sup> and H<sub>2</sub>O; (<b>e</b>) number of hydrogen bonds between ANI<sup>+</sup> and H<sub>2</sub>O influenced by simulated ultrasonication; (<b>f</b>) schematics of the activated ANI<sup>+</sup>.</p>
Full article ">Figure 6
<p>MD simulations of ANI<sup>+</sup> intercalating into the interlayer of bilayer graphene. (<b>a</b>–<b>c</b>) Snapshots of the aniline cations intercalating into the interlayer of bilayer graphene with a simulated time of (<b>a</b>) 0 ps, (<b>b</b>) 0.3 ps, and (<b>c</b>) 3 ps. Water molecules are hidden in the snapshots. The graphene is presented using paper chain model to exhibit its distortion. The intercalating aniline cations are presented using a van der Waals surface model (green). (<b>d</b>) RMSD curves of bilayer graphene intercalated by aniline cations with different initial kinetic energy; (<b>e</b>) relationships between the maximal RMSD values and the initial kinetic of aniline cations.</p>
Full article ">
30 pages, 8293 KiB  
Review
Biomimetic Nanomaterials: Diversity, Technology, and Biomedical Applications
by Kamil G. Gareev, Denis S. Grouzdev, Veronika V. Koziaeva, Nikita O. Sitkov, Huile Gao, Tatiana M. Zimina and Maxim Shevtsov
Nanomaterials 2022, 12(14), 2485; https://doi.org/10.3390/nano12142485 - 20 Jul 2022
Cited by 21 | Viewed by 5340
Abstract
Biomimetic nanomaterials (BNMs) are functional materials containing nanoscale components and having structural and technological similarities to natural (biogenic) prototypes. Despite the fact that biomimetic approaches in materials technology have been used since the second half of the 20th century, BNMs are still at [...] Read more.
Biomimetic nanomaterials (BNMs) are functional materials containing nanoscale components and having structural and technological similarities to natural (biogenic) prototypes. Despite the fact that biomimetic approaches in materials technology have been used since the second half of the 20th century, BNMs are still at the forefront of materials science. This review considered a general classification of such nanomaterials according to the characteristic features of natural analogues that are reproduced in the preparation of BNMs, including biomimetic structure, biomimetic synthesis, and the inclusion of biogenic components. BNMs containing magnetic, metal, or metal oxide organic and ceramic structural elements (including their various combinations) were considered separately. The BNMs under consideration were analyzed according to the declared areas of application, which included tooth and bone reconstruction, magnetic and infrared hyperthermia, chemo- and immunotherapy, the development of new drugs for targeted therapy, antibacterial and anti-inflammatory therapy, and bioimaging. In conclusion, the authors’ point of view is given about the prospects for the development of this scientific area associated with the use of native, genetically modified, or completely artificial phospholipid membranes, which allow combining the physicochemical and biological properties of biogenic prototypes with high biocompatibility, economic availability, and scalability of fully synthetic nanomaterials. Full article
(This article belongs to the Special Issue Biomimetic and Biogenic Multifunctional Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Biomimetic nanoparticles for biomedical applications. (<b>a</b>) Technology evolution: Early generations of particles were biologically inert and covered with nonfouling coatings, preventing their nonspecific interactions with the cells they encountered <span class="html-italic">in vivo</span>. From here, the next generation of nanoparticles became active, targeting molecules, which enabled the nanoparticles to reach the disease site and engage with the local environment. Taking inspiration from nature, the third generation of cell-membrane-based biomimetic nanoparticles mimicked the surface features of native cells by utilizing the whole cell membrane or membrane protein functionalization onto synthetic carriers (Reprinted from [<a href="#B31-nanomaterials-12-02485" class="html-bibr">31</a>], license CC BY 4.0.) (<b>b</b>) Schematic presentation of different strategies of inflammation-targeting biomimetic nanoparticles. Orange and red spheres represent drug-encapsulated synthetic nanoparticles (grey) and liposomes (green), respectively. (Reprinted from [<a href="#B16-nanomaterials-12-02485" class="html-bibr">16</a>], license CC BY 4.0.) (<b>c</b>) An example of modern BNM concept implementation: cell-membrane-coated NPs designed for atherosclerosis and inflammation therapy. The membranes of RBCs, platelets, and macrophages are extracted and used to coat different kinds of NPs depending on the site of inflammation and atherosclerosis. Each cell membrane has its own unique surface proteins, such as CD47 on RBC, integrin a4b1 on macrophages, and GPIIb/IIa on platelets, modifying the therapeutic effects. (Reprinted from [<a href="#B20-nanomaterials-12-02485" class="html-bibr">20</a>], license CC BY 4.0.)</p>
Full article ">Figure 1 Cont.
<p>Biomimetic nanoparticles for biomedical applications. (<b>a</b>) Technology evolution: Early generations of particles were biologically inert and covered with nonfouling coatings, preventing their nonspecific interactions with the cells they encountered <span class="html-italic">in vivo</span>. From here, the next generation of nanoparticles became active, targeting molecules, which enabled the nanoparticles to reach the disease site and engage with the local environment. Taking inspiration from nature, the third generation of cell-membrane-based biomimetic nanoparticles mimicked the surface features of native cells by utilizing the whole cell membrane or membrane protein functionalization onto synthetic carriers (Reprinted from [<a href="#B31-nanomaterials-12-02485" class="html-bibr">31</a>], license CC BY 4.0.) (<b>b</b>) Schematic presentation of different strategies of inflammation-targeting biomimetic nanoparticles. Orange and red spheres represent drug-encapsulated synthetic nanoparticles (grey) and liposomes (green), respectively. (Reprinted from [<a href="#B16-nanomaterials-12-02485" class="html-bibr">16</a>], license CC BY 4.0.) (<b>c</b>) An example of modern BNM concept implementation: cell-membrane-coated NPs designed for atherosclerosis and inflammation therapy. The membranes of RBCs, platelets, and macrophages are extracted and used to coat different kinds of NPs depending on the site of inflammation and atherosclerosis. Each cell membrane has its own unique surface proteins, such as CD47 on RBC, integrin a4b1 on macrophages, and GPIIb/IIa on platelets, modifying the therapeutic effects. (Reprinted from [<a href="#B20-nanomaterials-12-02485" class="html-bibr">20</a>], license CC BY 4.0.)</p>
Full article ">Figure 2
<p>Classification of biomimetic nanomaterials (BNMs) based on the literature data, including information on biomimetic structure [<a href="#B16-nanomaterials-12-02485" class="html-bibr">16</a>,<a href="#B34-nanomaterials-12-02485" class="html-bibr">34</a>,<a href="#B35-nanomaterials-12-02485" class="html-bibr">35</a>,<a href="#B36-nanomaterials-12-02485" class="html-bibr">36</a>,<a href="#B37-nanomaterials-12-02485" class="html-bibr">37</a>,<a href="#B38-nanomaterials-12-02485" class="html-bibr">38</a>], biomimetic synthesis [<a href="#B32-nanomaterials-12-02485" class="html-bibr">32</a>,<a href="#B39-nanomaterials-12-02485" class="html-bibr">39</a>,<a href="#B40-nanomaterials-12-02485" class="html-bibr">40</a>,<a href="#B41-nanomaterials-12-02485" class="html-bibr">41</a>,<a href="#B42-nanomaterials-12-02485" class="html-bibr">42</a>,<a href="#B43-nanomaterials-12-02485" class="html-bibr">43</a>], biogenic components [<a href="#B31-nanomaterials-12-02485" class="html-bibr">31</a>,<a href="#B33-nanomaterials-12-02485" class="html-bibr">33</a>,<a href="#B44-nanomaterials-12-02485" class="html-bibr">44</a>,<a href="#B45-nanomaterials-12-02485" class="html-bibr">45</a>,<a href="#B46-nanomaterials-12-02485" class="html-bibr">46</a>,<a href="#B47-nanomaterials-12-02485" class="html-bibr">47</a>], magnetic BNMs [<a href="#B48-nanomaterials-12-02485" class="html-bibr">48</a>,<a href="#B49-nanomaterials-12-02485" class="html-bibr">49</a>,<a href="#B50-nanomaterials-12-02485" class="html-bibr">50</a>,<a href="#B51-nanomaterials-12-02485" class="html-bibr">51</a>,<a href="#B52-nanomaterials-12-02485" class="html-bibr">52</a>,<a href="#B53-nanomaterials-12-02485" class="html-bibr">53</a>,<a href="#B54-nanomaterials-12-02485" class="html-bibr">54</a>], metal and metal oxide BNMs [<a href="#B55-nanomaterials-12-02485" class="html-bibr">55</a>,<a href="#B56-nanomaterials-12-02485" class="html-bibr">56</a>,<a href="#B57-nanomaterials-12-02485" class="html-bibr">57</a>,<a href="#B58-nanomaterials-12-02485" class="html-bibr">58</a>,<a href="#B59-nanomaterials-12-02485" class="html-bibr">59</a>,<a href="#B60-nanomaterials-12-02485" class="html-bibr">60</a>,<a href="#B61-nanomaterials-12-02485" class="html-bibr">61</a>], organic, ceramic and hybrid BNMs [<a href="#B62-nanomaterials-12-02485" class="html-bibr">62</a>,<a href="#B63-nanomaterials-12-02485" class="html-bibr">63</a>,<a href="#B64-nanomaterials-12-02485" class="html-bibr">64</a>,<a href="#B65-nanomaterials-12-02485" class="html-bibr">65</a>,<a href="#B66-nanomaterials-12-02485" class="html-bibr">66</a>,<a href="#B67-nanomaterials-12-02485" class="html-bibr">67</a>].</p>
Full article ">Figure 3
<p>Chitosan-oligosaccharide-coated biocompatible palladium nanoparticles (Pd@COS NPs) for photo-based imaging and therapy. (<b>a</b>) A scheme showing the preparation of Pd NPs, further surface coating with thiolated chitosan oligosaccharide (Pd@COS NPs), and finally, functionalization using an RGD peptide (Pd@COS-RGD). (<b>b</b>) A systematic illustration showing the photothermal ablation and photoacoustic imaging of tumor tissue using Pd@COS-RGD. (Reprinted from [<a href="#B69-nanomaterials-12-02485" class="html-bibr">69</a>], license CC BY 4.0.)</p>
Full article ">Figure 4
<p>Schematic illustration of encapsulation procedure and release mechanism of biocompatible upconversion nanoparticles (UCNP). (<b>a</b>) Encapsulation of UCNPs in a novel, synthesized phosphate surfactant through sonication at rt. (<b>b</b>) Release of UCNPs after a specific cleavage of phosphate surfactant by the sPLA-2 enzyme. (Reprinted from [<a href="#B75-nanomaterials-12-02485" class="html-bibr">75</a>], license CC BY 4.0.)</p>
Full article ">Figure 5
<p>The four main routes of the cytotoxic mechanism of AgNPs. 1: AgNPs adhere to the surface of a cell, damaging its membrane and altering the transport activity; 2: AgNPs and Ag ions penetrate inside the cell and interact with numerous cellular organelles and biomolecules, which can affect corresponding cellular functions; 3: AgNPs and Ag ions participate in the generation of reactive oxygen species (ROS) inside the cell, leading to cell damage; and 4: AgNPs and Ag ions induce the genotoxicity. (Reprinted from [<a href="#B81-nanomaterials-12-02485" class="html-bibr">81</a>], license CC BY 4.0.)</p>
Full article ">Figure 6
<p>HRTEM images of different NPs: (<b>a</b>,<b>b</b>) inorganic magnetite NPs, (<b>c</b>–<b>e</b>) MamC magnetite NPs, (<b>f</b>–<b>h</b>) Mms6 magnetite NPs, and (<b>i</b>–<b>k</b>) Mms6-MamC-mediated NPs. Selected areas of electron diffraction are shown for each sample. (Reprinted from [<a href="#B32-nanomaterials-12-02485" class="html-bibr">32</a>], license CC BY 4.0.)</p>
Full article ">Figure 7
<p>Magnetofection for gene delivery: (<b>A</b>) schematic representation of the process and (<b>B</b>) schematic illustration of DNA loading into lamellar magnetic hydroxyapatite (MHAp) nanoparticles for nucleic acid delivery. (Reprinted from [<a href="#B35-nanomaterials-12-02485" class="html-bibr">35</a>], license CC BY 3.0.)</p>
Full article ">Figure 8
<p>Surface features and scanning electron micrographs of a TiUnite dental implant surface. (Reprinted from [<a href="#B38-nanomaterials-12-02485" class="html-bibr">38</a>], license CC BY 4.0.)</p>
Full article ">Figure 9
<p>Schematic representation of the mechanism and final outcomes of the interaction of Au NPs and Ag NPs with a water dispersion of cubosomes and solid-supported films of cubosomes. (Reprinted from [<a href="#B34-nanomaterials-12-02485" class="html-bibr">34</a>], license CC BY 4.0.)</p>
Full article ">Figure 10
<p>Schematic illustration of biomimetically mineralized metal–organic framework (MOF). (<b>a</b>) Schematic of a sea urchin, a hard, porous, protective shell that is biomineralized by soft biological tissue. (<b>b</b>) Schematic of an MOF biocomposite showing a biomacromolecule (for example, protein, enzyme, or DNA) encapsulated within a porous, crystalline shell. (Reprinted from [<a href="#B108-nanomaterials-12-02485" class="html-bibr">108</a>], license CC BY 4.0.)</p>
Full article ">Figure 11
<p>Schematic illustration of mechanism of mitochondria-targeted cancer cell membrane biomimetic metal–organic framework mediated sonodynamic therapy and immune checkpoint blockade immunotherapy. (Reprinted from [<a href="#B115-nanomaterials-12-02485" class="html-bibr">115</a>], license CC BY 4.0.)</p>
Full article ">Figure 12
<p>Schematic of anisotropic nanoparticle fabrication and RBC membrane coating. (<b>A</b>) Spherical PLGA nanoparticles (NPs) were synthesized and cast onto a thin plastic film of 10% polyvinyl alcohol (PVA) and 2% glycerol. Particles were then stretched under heat in one and two dimensions (2D) to generate prolate and oblate ellipsoidal particles, respectively. (<b>B</b>) RBCs underwent hypotonic lysis and were then sonicated to generate sub—200 nm vesicles. RBC-derived vesicles were then coated on PLGA nanoparticles of all shapes under sonication. (Reprinted from [<a href="#B33-nanomaterials-12-02485" class="html-bibr">33</a>], license CC BY 4.0.)</p>
Full article ">Figure 13
<p>Plausible mechanism for the formation of Ag–TiO<sub>2</sub> NCs using <span class="html-italic">Origanum majorana</span> leaf extract. (Reprinted from [<a href="#B40-nanomaterials-12-02485" class="html-bibr">40</a>], license CC BY 4.0).</p>
Full article ">Figure 14
<p>Main applications of biomimetic nanomaterials.</p>
Full article ">
20 pages, 4344 KiB  
Article
Stable Dried Catalase Particles Prepared by Electrospraying
by Corinna S. Schlosser, Steve Brocchini and Gareth R. Williams
Nanomaterials 2022, 12(14), 2484; https://doi.org/10.3390/nano12142484 - 20 Jul 2022
Cited by 1 | Viewed by 2152
Abstract
Therapeutic proteins and peptides are clinically important, offering potency while reducing the potential for off-target effects. Research interest in developing therapeutic polypeptides has grown significantly during the last four decades. However, despite the growing research effort, maintaining the stability of polypeptides throughout their [...] Read more.
Therapeutic proteins and peptides are clinically important, offering potency while reducing the potential for off-target effects. Research interest in developing therapeutic polypeptides has grown significantly during the last four decades. However, despite the growing research effort, maintaining the stability of polypeptides throughout their life cycle remains a challenge. Electrohydrodynamic (EHD) techniques have been widely explored for encapsulation and delivery of many biopharmaceuticals. In this work, we explored monoaxial electrospraying for encapsulation of bovine liver catalase, investigating the effects of the different components of the electrospraying solution on the integrity and bioactivity of the enzyme. The catalase was successfully encapsulated within polymeric particles made of polyvinylpyrrolidone (PVP), dextran, and polysucrose. The polysorbate 20 content within the electrospraying solution (50 mM citrate buffer, pH 5.4) affected the catalase loading—increasing the polysorbate 20 concentration to 500 μg/mL resulted in full protein encapsulation but did not prevent loss in activity. The addition of ethanol (20% v/v) to a fully aqueous solution improves the electrospraying process by reducing surface tension, without loss of catalase activity. The polymer type was shown to have the greatest impact on preserving catalase activity within the electrosprayed particles. When PVP was the carrier there was no loss in activity compared with fresh aqueous solutions of catalase. The optimum particles were obtained from a 20% w/v PVP or 30% w/v PVP-trehalose (1:1 w/w) solution. The addition of trehalose confers stability advantages to the catalase particles. When trehalose-PVP particles were stored at 5 °C, enzymatic activity was maintained over 3 months, whereas for the PVP-only analogue a 50% reduction in activity was seen. This demonstrates that processing catalase by monoaxial electrospraying can, under optimised conditions, result in stable polymeric particles with no loss of activity. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Area under the curve of catalase solutions (50 μg/mL), determined by SEC after storage at different ethanol concentrations (<b>a</b>) peak 1, (<b>b</b>) peak 2.</p>
Full article ">Figure 2
<p>Catalase activity after storage at pH 5.4 in the presence of polysorbate 20 or polysorbate 80 at three different surfactant-to-catalase ratios (<b>a</b>) at 50 μg/mL catalase for 6 h, (<b>b</b>) at 1 mg/mL catalase for 24 h.</p>
Full article ">Figure 3
<p>Protein content and activity for the catalase-dextran (1 μg/mg) particles electrosprayed from solutions containing 0–500 μg/mL of polysorbate 20. (<b>a</b>) catalase content, (<b>b</b>) catalase activity.</p>
Full article ">Figure 4
<p>Protein content and activity for catalase-polymer particles (1 μg/mg) containing three different polymers (10% <span class="html-italic">w</span>/<span class="html-italic">v</span> in solution): (<b>a</b>) catalase content, (<b>b</b>) catalase activity.</p>
Full article ">Figure 5
<p>Protein content and activity in electrosprayed catalase-PVP (1 μg/mg) and catalase-PVP-trehalose (1 μg/mg) particles obtained from 10%, 15%, and 20% <span class="html-italic">w</span>/<span class="html-italic">v</span> PVP or 20% <span class="html-italic">w</span>/<span class="html-italic">v</span> and 30% PVP-trehalose (1:1) solutions.</p>
Full article ">Figure 6
<p>Scanning electron microscopy images of the electrosprayed catalase-PVP (1 μg/mg) particles prepared with solutions containing (<span class="html-italic">w</span>/<span class="html-italic">v</span>) (<b>a</b>) 10% PVP, (<b>b</b>) 15% PVP, (<b>c</b>) 20% PVP, (<b>d</b>) 20% PVP-trehalose, and (<b>e</b>) 30% PVP-trehalose. The scale bar represents 10 μm.</p>
Full article ">Figure 7
<p>Particle size distribution of the electrosprayed catalase-PVP particles prepared from solutions containing (<span class="html-italic">w</span>/<span class="html-italic">v</span>) (<b>a</b>) 10% PVP, (<b>b</b>) 15% PVP, (<b>c</b>) 20% PVP, (<b>d</b>) 20% PVP-trehalose, and (<b>e</b>) 30% PVP-trehalose.</p>
Full article ">Figure 8
<p>FTIR spectra of the raw materials and protein-loaded particles prepared from PVP solutions, showing (<b>a</b>) full spectrum and (<b>b</b>) enlargement of the 1750–650 cm<sup>−1</sup> region, and PVP-trehalose solutions: (<b>c</b>) full spectrum and (<b>d</b>) enlargement of the 1750–650 cm<sup>−1</sup> region.</p>
Full article ">Figure 8 Cont.
<p>FTIR spectra of the raw materials and protein-loaded particles prepared from PVP solutions, showing (<b>a</b>) full spectrum and (<b>b</b>) enlargement of the 1750–650 cm<sup>−1</sup> region, and PVP-trehalose solutions: (<b>c</b>) full spectrum and (<b>d</b>) enlargement of the 1750–650 cm<sup>−1</sup> region.</p>
Full article ">Figure 9
<p>XRD diffraction patterns of the formulations and raw materials. (<b>a</b>) PVP particles, (<b>b</b>) PVP-trehalose particles.</p>
Full article ">Figure 10
<p>Residual activity based on measured protein content within the polymeric particles (1 μg/mg catalase:polymer) and unformulated catalase stored over 90 days at (<b>a</b>) 5 °C, (<b>b</b>) room temperature, and (<b>c</b>) 40 °C/75% RH. Where the 30% PVP-trehalose formulation is statistically significant different to either 20% PVP (dark blue asterisk) or unprocessed catalase (light blue asterisk) at a given time point, this has been indicated by (*) <span class="html-italic">p</span> &lt; 0.05 and (**) <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
9 pages, 2825 KiB  
Article
Study on Nanoporous Graphene-Based Hybrid Architecture for Surface Bonding
by Xiaohui Song, Mingxiang Chen, Jingshuang Zhang, Rui Zhang and Wei Zhang
Nanomaterials 2022, 12(14), 2483; https://doi.org/10.3390/nano12142483 - 20 Jul 2022
Cited by 4 | Viewed by 1637
Abstract
Graphene-copper nanolayered composites have received research interest as promising packaging materials in developing next-generation electronic and optoelectronic devices. The weak van der Waal (vdW) contact between graphene and metal matrix significantly reduces the mechanical performance of such composites. The current study describes a [...] Read more.
Graphene-copper nanolayered composites have received research interest as promising packaging materials in developing next-generation electronic and optoelectronic devices. The weak van der Waal (vdW) contact between graphene and metal matrix significantly reduces the mechanical performance of such composites. The current study describes a new Cu-nanoporous graphene-Cu based bonding method with a low bonding temperature and good dependability. The deposition of copper atoms onto nanoporous graphene can help to generate nanoislands on the graphene surface, facilitating atomic diffusion bonding to bulk copper bonding surfaces at low temperatures, according to our extensive molecular dynamics (MD) simulations on the bonding process and pull-out verification using the canonical ensemble (NVT). Furthermore, the interfacial mechanical characteristics of graphene/Cu nanocomposites can be greatly improved by the resistance of nanostructure in nanoporous graphene. These findings are useful in designing advanced metallic surface bonding processes and graphene-based composites with tenable performance. Full article
(This article belongs to the Special Issue Mechanics of Micro- and Nano-Size Materials and Structures)
Show Figures

Figure 1

Figure 1
<p>Atomic configurations. (<b>a</b>) Cu-nanoporous graphene composite model; (<b>b</b>) Cu atoms deposition onto the nanoporous graphene surface; (<b>c</b>) Cu-nanoporous graphene-Cu thermocompression bonding; (<b>d</b>) Pull-out simulation model.</p>
Full article ">Figure 2
<p>Atomic configurations of Cu atoms deposition onto the graphene at various stages. (<b>a</b>) Filling of nanopores at the initial stage; (<b>b</b>) Insular growth to nanoislands; (<b>c</b>) Further growing of nanoislands; (<b>d</b>) Joining with neighboring nanoislands.</p>
Full article ">Figure 3
<p>Cross-section configurations of Cu-nanoporous graphene-Cu during the bonding process at various stages. (<b>a</b>) Contacting of bonding surface; (<b>b</b>) Compression deformation of the nanoislands; (<b>c</b>) Final bonding structure.</p>
Full article ">Figure 4
<p>The pull-out force changes as a function of the sliding distance for the graphene/Cu composite with and without nanoporous.</p>
Full article ">Figure 5
<p>The displacement of copper atoms at the interface of graphene/Cu nanocomposites with various sliding distances.</p>
Full article ">Figure 6
<p>The maximum pull-out force changes as a function of the number of deposited copper atoms for the graphene/Cu composite.</p>
Full article ">Figure 7
<p>The maximum pull-out force changes vs. nanoporous diameters.</p>
Full article ">
13 pages, 4848 KiB  
Article
Covalent Organic Framework/Polyacrylonitrile Electrospun Nanofiber for Dispersive Solid-Phase Extraction of Trace Quinolones in Food Samples
by Jinghui Zhou, An Chen, Hongying Guo, Yijun Li, Xiwen He, Langxing Chen and Yukui Zhang
Nanomaterials 2022, 12(14), 2482; https://doi.org/10.3390/nano12142482 - 20 Jul 2022
Cited by 14 | Viewed by 2425
Abstract
The extraction of quinolone antibiotics (QAs) is crucial for the environment and human health. In this work, polyacrylonitrile (PAN)/covalent organic framework TpPa–1 nanofiber was prepared by an electrospinning technique and used as an adsorbent for dispersive solid-phase extraction (dSPE) of five QAs in [...] Read more.
The extraction of quinolone antibiotics (QAs) is crucial for the environment and human health. In this work, polyacrylonitrile (PAN)/covalent organic framework TpPa–1 nanofiber was prepared by an electrospinning technique and used as an adsorbent for dispersive solid-phase extraction (dSPE) of five QAs in the honey and pork. The morphology and structure of the adsorbent were characterized, and the extraction and desorption conditions for the targeted analytes were optimized. Under the optimal conditions, a sensitive method was developed by using PAN/TpPa–1 nanofiber as an adsorbent coupled with high-performance liquid chromatography (HPLC) for five QAs detection. It offered good linearity in the ranges of 0.5–200 ng·mL−1 for pefloxacin, enrofloxacin, and orbifloxacin, and of 1–200 ng·mL−1 for norfloxacin and sarafloxacin with correlation coefficients above 0.9946. The limits of detection (S/N = 3) of five QAs ranged from 0.03 to 0.133 ng·mL−1. The intra-day and inter-day relative standard deviations of the five QAs with the spiked concentration of 50 ng·mL−1 were 2.8–4.0 and 3.0–8.8, respectively. The recoveries of five QAs in the honey and pork samples were 81.6–119.7%, which proved that the proposed method has great potential for the efficient extraction and determination of QAs in complex samples. Full article
(This article belongs to the Special Issue Nanomaterials-Based Sample Pretreatment)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the preparation of PAN/TpPa–1 nanofiber and used as a dSPE adsorbent of QAs.</p>
Full article ">Figure 2
<p>SEM images of TpPa–1 (<b>A</b>), pure PAN (<b>B</b>), PAN/TpPa–1 (<b>C</b>,<b>D</b>) nanofiber.</p>
Full article ">Figure 3
<p>FT-IR spectra of Tp, Pa–1, and COF TpPa–1.</p>
Full article ">Figure 4
<p>XRD patterns of TpPa–1, PAN/TpPa–1 nanofiber, pure PAN nanofiber, and simulation of TpPa–1.</p>
Full article ">Figure 5
<p>Effect of amount of PAN/TpPa–1 nanofiber (<b>A</b>), types of desorption solvent (<b>B</b>), desorption time (<b>C</b>), extraction time (<b>D</b>), concentration of NaCl (<b>E</b>), and standard solution pH value (<b>F</b>) on the extraction performance of QAs.</p>
Full article ">Figure 6
<p>(<b>A</b>): HPLC-UV chromatograms of (a) honey sample; (b) honey sample after dSPE; (c) honey sample spiked with 100 ng·mL<sup>–1</sup> before dSPE; and (d) honey sample spiked 100 ng·mL<sup>–1</sup> after dSPE. (<b>B</b>): HPLC-UV chromatograms of (a) pork sample; (b) pork sample after dSPE; (c) pork sample spiked with 100 ng·mL<sup>–1</sup> before dSPE; and (d) pork sample spiked 100 ng·mL<sup>–1</sup> after dSPE.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop