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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (334)

Search Parameters:
Keywords = atomizing sheet

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 882 KiB  
Article
A DFT Study of the Mechanical Properties of a Lizardite Slab Reinforced by Graphene and Hexagonal Boron Nitride
by Anne Karollynne Castro Monteiro, Consuelo Alves da Frota, Cicero Mota, Angsula Ghosh and Hidembergue Ordozgoith da Frota
Minerals 2025, 15(1), 53; https://doi.org/10.3390/min15010053 - 7 Jan 2025
Viewed by 357
Abstract
The stacking of two-dimensional atomic-level thickness materials onto hexagonal boron nitride (h-BN) and graphene (Gr) not only significantly enhances their properties, but also exhibits a multitude of exceptional characteristics, promising widespread applications across various fields. Clay minerals hold profound significance in scientific research [...] Read more.
The stacking of two-dimensional atomic-level thickness materials onto hexagonal boron nitride (h-BN) and graphene (Gr) not only significantly enhances their properties, but also exhibits a multitude of exceptional characteristics, promising widespread applications across various fields. Clay minerals hold profound significance in scientific research not only because of their abundance but also because of their application in geology, environmental science, materials science, and biotechnology. We present a study that uses density functional theory (DFT) to analyze the effect on the mechanical properties of lizardite slab-reinforced Gr or h-BN monolayers. In addition to the reference lizardite slab (Liza-2D), six composites were studied: a monolayer of Gr (h-BN) over the octahedral face of a pristine lizardite slab (Liza-Gr1 (Liza-BN1)), a monolayer of Gr (h-BN) under the tetrahedral face of a pristine lizardite slab (Liza-Gr2(Liza-BN2)), and a pristine lizardite slab sandwiched between two Gr (h-BN) monolayers (Liza-Gr3(Liza-BN3)). We observed that reinforcement by Gr or h-BN significantly increased the bulk, Young’s and shear moduli of the composites. Taking into account that the Gr and h-BN sheets interact weakly by van der Waals interactions with the lizardite slab surface, we estimated the Young’s and shear moduli of the composites by the Rule of Mixtures and obtained a reasonable agreement with those from DFT calculations. Full article
(This article belongs to the Section Clays and Engineered Mineral Materials)
Show Figures

Figure 1

Figure 1
<p>The superstructures to study the contribution of graphene and h-BN for the reinforcement of lizardite: (<b>a</b>) Liza—a slab extracted from the pristine lizardite layered crystal; (<b>b</b>,<b>e</b>) Liza-Gr1 (Liza-BN1)—a monolayer of Gr (h-BN) over the octahedral face of the pristine lizardite slab; (<b>c</b>,<b>f</b>) Liza-Gr2 (Liza-BN2)—a monolayer of Gr (h-BN) under the tetrahedral face of the pristine lizardite slab; (<b>d</b>,<b>g</b>) Liza-Gr3 (Liza-BN3)—a slab of pristine lizardite sandwiched between two monolayers of Gr (h-BN).</p>
Full article ">Figure 2
<p>Comparison of the bulk, Young’s and shear moduli of the composites with the reference Liza-Gr: (<b>a</b>) Liza-Gr1, Liza-Gr2 and Liza-Gr3; (<b>b</b>) Liza-BN1, Liza-BN2 and Liza-BN3. The reinforcement by both Gr and h-BN sheets significantly increases the mechanical properties of lizardite slab.</p>
Full article ">Figure 3
<p>Comparison of the Young’s and shear moduli of the composites obtained by the DFT method and estimated by the Rule of Mixtures (ROM): (<b>a</b>) lizardite slab reinforced by Gr; (<b>b</b>) lizardite slab reinforced by h-BN. In both figures, Y(DFT) and G(DFT) (Y(ROM) and G(ROM)) represent the Young’s and shear moduli obtained by DFT (estimated by ROM).</p>
Full article ">
9 pages, 2942 KiB  
Communication
Crystal Structure of the Biocide Methylisothiazolinone
by Richard Goddard, Rüdiger W. Seidel, Michael Patzer and Nils Nöthling
Crystals 2024, 14(12), 1100; https://doi.org/10.3390/cryst14121100 - 20 Dec 2024
Viewed by 427
Abstract
Methylisothiazolinone (MIT) is widely used as a biocide in numerous personal care products, glass-cleaning liquids, paints, and industrial applications. MIT and related isothiazolinones have attracted much attention for their allergenic properties such as contact dermatitis. Although the compound was first prepared in 1964 [...] Read more.
Methylisothiazolinone (MIT) is widely used as a biocide in numerous personal care products, glass-cleaning liquids, paints, and industrial applications. MIT and related isothiazolinones have attracted much attention for their allergenic properties such as contact dermatitis. Although the compound was first prepared in 1964 and has been widely used as a biocide since the 1970s, its crystal structure has so far not been reported. Here we report the solid state structure of MIT as determined by single crystal X-ray diffraction (SC-XRD) analysis of a crystal grown from the melt. MIT crystallizes as a layered structure with short C-H···O hydrogen bonding interactions within the sheets. The average distance between the sheets parallel to (1 0 2) is ca. 3.2 Å. The molecule exhibits a small C-S-N angle of 90.81(2)° and a methyl group that is slightly bent out of the plane of the planar five-membered ring. The sulfur atom does not undergo any significant intermolecular interactions. Full article
(This article belongs to the Section Crystal Engineering)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Molecular structure of MIT in the crystal. Displacement ellipsoids are drawn at the 50% probability level. (<b>b</b>) Packing of the two molecules of MIT in the triclinic unit cell. Hydrogen atoms are omitted for clarity. o, a, b, and c, denote the origin and unit cell axes, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) Depiction of the layered arrangement of the molecules in the crystal of MIT. (<b>b</b>) View approximately perpendicular to the sheet highlighting the close packing of the molecules. Hydrogen atoms have been omitted for clarity. 0, a, b, and c, denote the origin and unit cell axes, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) Hirshfeld surface mapped with <span class="html-italic">d</span><sub>norm</sub> showing the major intermolecular interactions within the crystalline structure of MIT; (<b>b</b>) 2D fingerprint plot of the Hirshfeld surface providing an overview of the intermolecular interactions.</p>
Full article ">Figure 4
<p>Superposition of the N-S-C units of the experimentally determined geometry of MIT (green) with those of the DFT optimized structure of the free molecule (pink) and MCI in the crystal structure (blue, CSD refcode: XIFRIO), illustrating the non-planarity of the nitrogen atom in the crystal structures of MIT and MCI and the different arrangements of the methyl hydrogen atoms.</p>
Full article ">Scheme 1
<p>Structural formula of 2-methyl-2<span class="html-italic">H</span>-isothiazol-3-one (MIT).</p>
Full article ">
14 pages, 3223 KiB  
Article
Molecular Conductors Based on Dimethylcyclohexene-Fused Tetrathiafulvalene
by Masahiro Fujisaki, Ryoya Naito, Takashi Shirahata, Yoshitaka Kawasugi, Naoya Tajima and Yohji Misaki
Chemistry 2024, 6(6), 1509-1522; https://doi.org/10.3390/chemistry6060091 - 25 Nov 2024
Viewed by 730
Abstract
Chiral electroactive materials have attracted attention for the effects of electrical magnetochiral anisotropy (eMChA) and chirality-induced spin selectivity (CISS). The combination of tetrathiafulvalene (TTF) with chiral moieties is one way to access chiral electroactive materials. In this paper, we have focused on the [...] Read more.
Chiral electroactive materials have attracted attention for the effects of electrical magnetochiral anisotropy (eMChA) and chirality-induced spin selectivity (CISS). The combination of tetrathiafulvalene (TTF) with chiral moieties is one way to access chiral electroactive materials. In this paper, we have focused on the fused 2,3-dimethylcyclohexene (DMCh) ring as a substituent with chiral carbon atoms and without heteroatoms, which has not been used in the field of molecular conductors, and we synthesized a new TTF derivative (rac-DMCh-EDT-TTF). We have developed novel molecular conductors (rac-DMCh-EDT-TTF)2X (X = PF6, AsF6 and ClO4), which have bilayer conducting sheets composed of the two crystallographically independent molecules. All salts exhibited semiconducting behavior from room temperature down to low temperatures, and a resistivity anomaly was observed at 180–250 K. X-ray structure analysis at 100 K and 263 K and molecular orbital calculations using the results of X-ray structure analysis indicated the emergence of a charge disproportionation between Layers 1 and 2 at the low-temperature phase. Full article
(This article belongs to the Section Inorganic and Solid State Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The temperature dependence of the electrical resistivity of the <span class="html-italic">rac</span>-DMCh-EDT-TTF salts. (<b>b</b>) The temperature dependence of the electrical resistivity, plotted as log (<span class="html-italic">ρ</span>) versus 1/<span class="html-italic">T</span>, for <span class="html-italic">rac</span>-DMCh-EDT-TTF salts at ambient pressure.</p>
Full article ">Figure 2
<p>(<b>a</b>) The temperature dependence of the electrical resistivity of the (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>ClO<sub>4</sub> under pressures up to 1.8 GPa. (<b>b</b>) The temperature dependence of the electrical resistivity, plotted as log (<span class="html-italic">ρ</span>) versus 1/<span class="html-italic">T</span>, for (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>ClO<sub>4</sub> under pressures up to 1.8 GPa.</p>
Full article ">Figure 3
<p>The crystal structure of (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>PF<sub>6</sub> at 100 K, (<b>a</b>) viewed along the <span class="html-italic">a</span> axis and (<b>b</b>) viewed along the <span class="html-italic">b</span> axis. Dash lines represent C-F contacts (Red: C21-F5 = 3.100(3) Å, Blue: C28-F4 = 3.157(3) Å). The conducting sheet viewed along the donor long axis of (<b>c</b>) Layer 1 and (<b>d</b>) Layer 2 in (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>PF<sub>6</sub> at 100 K. Hydrogen atoms are omitted for clarity.</p>
Full article ">Figure 4
<p>The HOMO (isocontour value = 0.03 a.u.) of donor molecule 1 (<b>a</b>,<b>b</b>) and donor molecule 2 (<b>c</b>,<b>d</b>) in (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>PF<sub>6</sub>. The calculations were performed by the extended Hückel method based on the molecular structures determined by crystal structure analyses of (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>PF<sub>6</sub> at 100 K.</p>
Full article ">Figure 5
<p>The calculated density of states (DOS), band dispersion, and Fermi surfaces of (<span class="html-italic">rac</span>-DMCh-EDT-TTF)<sub>2</sub>PF<sub>6</sub> at (<b>a</b>) 263 K and (<b>b</b>) 100 K. The band dispersion and Fermi surface are represented by the red lines for Layer 1 and the blue lines for Layer 2.</p>
Full article ">Scheme 1
<p>The synthesis of <span class="html-italic">rac</span>-DMCh-EDT-TTF.</p>
Full article ">
24 pages, 7292 KiB  
Article
The Impact of Temperature and Pressure on the Structural Stability of Solvated Solid-State Conformations of Bombyx mori Silk Fibroins: Insights from Molecular Dynamics Simulations
by Ezekiel Edward Nettey-Oppong, Riaz Muhammad, Ahmed Ali, Hyun-Woo Jeong, Young-Seek Seok, Seong-Wan Kim and Seung Ho Choi
Materials 2024, 17(23), 5686; https://doi.org/10.3390/ma17235686 - 21 Nov 2024
Viewed by 848
Abstract
Bombyx mori silk fibroin is a promising biopolymer with notable mechanical strength, biocompatibility, and potential for diverse biomedical applications, such as tissue engineering scaffolds, and drug delivery. These properties are intrinsically linked to the structural characteristics of silk fibroin, making it essential to [...] Read more.
Bombyx mori silk fibroin is a promising biopolymer with notable mechanical strength, biocompatibility, and potential for diverse biomedical applications, such as tissue engineering scaffolds, and drug delivery. These properties are intrinsically linked to the structural characteristics of silk fibroin, making it essential to understand its molecular stability under varying environmental conditions. This study employed molecular dynamics simulations to examine the structural stability of silk I and silk II conformations of silk fibroin under changes in temperature (298 K to 378 K) and pressure (0.1 MPa to 700 MPa). Key parameters, including Root Mean Square Deviation (RMSD), Root Mean Square Fluctuation (RMSF), and Radius of Gyration (Rg) were analyzed, along with non-bonded interactions such as van der Waals and electrostatic potential energy. Our findings demonstrate that both temperature and pressure exert a destabilizing effect on silk fibroin, with silk I exhibiting a higher susceptibility to destabilization compared to silk II. Additionally, pressure elevated the van der Waals energy in silk I, while temperature led to a reduction. In contrast, electrostatic potential energy remained unaffected by these environmental conditions, highlighting stable long-range interactions throughout the study. Silk II’s tightly packed β-sheet structure offers greater resilience to environmental changes, while the more flexible α-helices in silk I make it more susceptible to structural perturbations. These findings provide valuable insights into the atomic-level behavior of silk fibroin, contributing to a deeper understanding of its potential for applications in environments where mechanical or thermal stress is a factor. The study underscores the importance of computational approaches in exploring protein stability and supports the continued development of silk fibroin for biomedical and engineering applications. Full article
(This article belongs to the Special Issue Advances in Bio-Polymer and Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of <span class="html-italic">Bombyx mori</span> silk structure. During the pupa stage of their metamorphosis into moths, silkworms spin silk fibers to construct protective cocoons. Each silk fiber is composed of two core fibroin filaments encased by sericin (depicted in purple), an adhesive glycoprotein that facilitates fiber cohesion. At the molecular level, each fibroin filament consists of numerous assemblies of nanofibrils, which can adopt either silk I or silk II structural conformations. These structural forms are determined by the specific arrangement of secondary protein structures within the fibroin. The silk I structure predominantly features type II β-turns and α-helices, while the silk II structure is mainly characterized by β-turns and β-sheets. Both of these secondary structures contribute to the fiber’s mechanical properties and functional versatility across various applications.</p>
Full article ">Figure 2
<p>Schematic of the secondary structures of <span class="html-italic">Bombyx mori</span> silk fibroin. The silk fibroin protein comprises several distinct secondary structures that play a critical role in determining the material’s mechanical strength and biological properties. These secondary structures include the following: (<b>a</b>) α-helix, a right-handed coiled structure that contributes to flexibility; (<b>b</b>) β-sheet, an extended conformation that forms the crystalline regions, providing mechanical robustness; and (<b>c</b>) random coils, which are unstructured regions that contribute to the amorphous domains of the protein. The schematic also illustrates the specific amino acid residues associated with each of these secondary structures, alongside the corresponding all-atom models, providing a molecular-level perspective on silk fibroin’s hierarchical organization.</p>
Full article ">Figure 3
<p>Schematic representation of the primary structure of <span class="html-italic">Bombyx mori</span> silk fibroin. The heavy chain of silk fibroin consists of alternating crystalline (R1 to R12) and amorphous phases (L1 to L11), along with N terminus (N) and C terminus (C), each having distinct structural and functional roles. The crystalline regions, indicated by the R domains, are primarily composed of β-sheet structures that impart rigidity and strength to the fiber. In contrast, the amorphous regions, denoted as L domains, serve as flexible linkers that provide elasticity and contribute to the material’s overall mechanical performance. This schematic highlights the representative molecular structure used for simulations, where domain R6 (crystalline) is flanked by domains L5 and L6 (amorphous), providing a balanced model for studying both phases of the fibroin.</p>
Full article ">Figure 4
<p>Molecular dynamics equilibration simulation of silk fibroin protein structures. This figure illustrates the evolution of both potential energy and kinetic energy over time during the equilibration phase for hydrated silk fibroin systems. Equilibration was performed under the NPT ensemble for the silk I (<b>a</b>) and silk II (<b>b</b>) structures. The gradual decrease and stabilization of the total potential energy throughout the simulation indicates that the system achieves a stable configuration as the atoms settle into their equilibrium positions. The kinetic energy, in contrast, remains relatively constant, reflecting consistent thermal motion within the system. This energy evolution demonstrates the successful stabilization of silk fibroin structures under the specified conditions, preparing them for further simulation analyses.</p>
Full article ">Figure 5
<p>Volume changes of the simulation cells during equilibration of silk fibroin structures. The figure presents the observed reduction in the volume of the simulation cells during equilibration. For the silk I structure (<b>a</b>), the cubic cell size decreased from an initial length of 105 Å to 102.7 Å, while for the silk II structure (<b>b</b>), the cell size reduced from 89 Å to 86.9 Å. This volume contraction is indicative of the system reaching equilibrium, as the protein and water molecules reorganize into a more compact and energetically favorable configuration. The simulations ensured complete hydration of the protein by maintaining a 10 Å buffer between the protein structures and the cell boundaries. A visual snapshot on the far right provides a cross-sectional view of the hydrated silk I and silk II proteins, illustrating the distribution of water molecules around the protein structures and confirming that the proteins are fully solvated within the simulation cells.</p>
Full article ">Figure 6
<p>Root Mean Square Deviation (RMSD) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) RMSD values for the backbone atoms of silk I across a pressure range of 0.1 MPa to 700 MPa, showing how the structural deviation increases with pressure. (<b>b</b>) RMSD values for the backbone atoms of silk II under the same pressure range. Both structures exhibit increasing deviation with pressure. (<b>c</b>) The average RMSD values for silk I and silk II as a function of pressure. Silk I has a minimum deviation of 0.804 Å at 0.1 MPa, increasing to a maximum of 1.454 Å at 700 MPa. Similarly, silk II exhibits a minimum deviation of 0.772 Å at 0.1 MPa and a maximum of 1.285 Å at 700 MPa. These results demonstrate that pressure induces structural instability, with silk I experiencing a more pronounced deviation than silk II. (<b>d</b>) RMSD of silk I backbone atoms across a temperature range of 298 K to 378 K, indicating how thermal agitation affects structural deviation. (<b>e</b>) RMSD values for silk II at the same temperature range, illustrating the temperature-dependent structural perturbations. (<b>f</b>) The average RMSD values for silk I and silk II as a function of temperature. The temperature increase caused a rise in RMSD for both structures, with silk I showing a deviation from 0.764 Å at 298 K to 0.871 Å at 378 K, and silk II deviating from 0.742 Å to 0.843 Å over the same temperature range. This increase indicates that thermal agitation leads to greater atomic movement and structural deviations, particularly for silk I.</p>
Full article ">Figure 7
<p>Root Mean Square Fluctuation (RMSF) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) RMSF values for silk I backbone atoms across a pressure range of 0.1 MPa to 700 MPa, showing per-residue fluctuations and the effect of pressure on protein flexibility. (<b>b</b>) RMSF values for silk II under the same pressure range, highlighting differences in flexibility between the two structures. (<b>c</b>) The average RMSF values for silk I and silk II as a function of pressure. The minimum fluctuation for silk I was 0.470 Å at 0.1 MPa, increasing to a maximum of 0.507 Å at 700 MPa. For silk II, the fluctuation values ranged from 0.454 Å to 0.453 Å over the same pressure range. The observed fluctuations are relatively low, indicating minimal atomic mobility under pressure for both structures. (<b>d</b>) RMSF values for silk I backbone atoms across a temperature range of 298 K to 378 K, showing increased fluctuations with rising temperature. (<b>e</b>) RMSF values for silk II at the same temperature range, reflecting a similar trend of increased fluctuations with temperature. (<b>f</b>) The average RMSF values for silk I and silk II as a function of temperature. The fluctuations increase with temperature, with silk I showing a rise from 0.498 Å at 298 K to 0.592 Å at 378 K, and silk II increasing from 0.468 Å to 0.536 Å. These results suggest that temperature-induced thermal agitation leads to increased per-residue flexibility, particularly for silk I, which demonstrates greater fluctuation than silk II.</p>
Full article ">Figure 8
<p>Radius of Gyration (R<sub>g</sub>) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) R<sub>g</sub> values for silk I across a pressure range of 0.1 MPa to 700 MPa, illustrating how pressure affects the overall compactness of the structure. (<b>b</b>) R<sub>g</sub> values for silk II under the same pressure range, showing a similar trend but with lower values compared to silk I. (<b>c</b>) The average R<sub>g</sub> values for silk I and silk II as a function of pressure. An increase in pressure leads to a reduction in R<sub>g</sub>, indicating increased compaction of both structures. Silk I shows a maximum R<sub>g</sub> of 22.485 Å at 0.1 MPa, decreasing to 21.520 Å at 700 MPa. Silk II exhibits a maximum R<sub>g</sub> of 20.635 Å at 0.1 MPa, reducing to 19.929 Å at 700 MPa. These results confirm that pressure induces compaction, with silk I showing a greater reduction in compactness than silk II. (<b>d</b>) R<sub>g</sub> values for silk I across a temperature range of 298 K to 378 K, indicating how thermal effects impact protein packing. (<b>e</b>) R<sub>g</sub> values for silk II under the same temperature range, showing minimal changes in compactness. (<b>f</b>) The average R<sub>g</sub> values for silk I and silk II as a function of temperature. Both structures show minimal alterations in compactness with temperature, with R<sub>g</sub> values of 22.762 Å for silk I and 20.874 Å for silk II at 298 K, slightly decreasing to 22.711 Å and 20.937 Å, respectively, at 378 K. The small changes in R<sub>g</sub> suggest that the temperature range studied has a negligible effect on the compactness of the silk structures.</p>
Full article ">Figure 9
<p>Non-bonded interactions of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) The average van der Waals energy as a function of pressure, from 0.1 MPa to 700 MPa. For both silk I and silk II, the van der Waals energy increased with rising pressure, reflecting the compression of inter-atomic distances and the strengthening of long-range non-bonded interactions. (<b>b</b>) The average electrostatic potential energy as a function of pressure. The electrostatic potential energy remained constant for both silk I and silk II, with values of −1.97 × 10<sup>6</sup> kcal/mol and −1.16 × 10<sup>6</sup> kcal/mol, respectively, indicating that pressure has no significant effect on electrostatic interactions. (<b>c</b>) The average van der Waals energy as a function of temperature, from 298 K to 378 K. In contrast to pressure, the van der Waals energy decreased with increasing temperature for both silk structures, due to the expansion of inter-atomic distances and the weakening of non-bonded interactions. (<b>d</b>) The average electrostatic potential energy as a function of temperature. Similar to pressure, the electrostatic potential energy remained constant with temperature for both silk I and silk II, highlighting the stability of long-range electrostatic interactions under thermal fluctuations.</p>
Full article ">
11 pages, 2737 KiB  
Article
Verification of Optimal X-Ray Shielding Properties Based on Material Composition and Coating Design of Shielding Materials
by Seon-Chil Kim, Jae-Han Yun, Hong-Sik Byun and Jian Hou
Coatings 2024, 14(11), 1450; https://doi.org/10.3390/coatings14111450 - 14 Nov 2024
Viewed by 820
Abstract
Health care workers performing radiography on patients in hospitals typically wear aprons for radiation protection. Protective properties are achieved through a combination of shielding materials and polymers. Various shielding materials are mixed with polymers to prepare composite materials. Numerous methods have been devised [...] Read more.
Health care workers performing radiography on patients in hospitals typically wear aprons for radiation protection. Protective properties are achieved through a combination of shielding materials and polymers. Various shielding materials are mixed with polymers to prepare composite materials. Numerous methods have been devised to design and alter the composition of these materials to improve the shielding performance of aprons. In this study, the shielding performance was analyzed based on the arrangement of shielding materials, the composition of materials (mixed or single), and the fabrication design of the shielding sheets. Various shielding sheets were created using different arrangements of tungsten oxide, bismuth oxide, and barium sulfate, and their shielding efficacy was compared. The atomic number and density of the shielding material directly affect the shielding property. The effectiveness of the composite sheet increased by more than 5% when positioned close to the X-ray tube. Sheets fabricated from materials separated by type, rather than mixed, exhibited a greater X-ray shielding effectiveness because of their layered structure. Therefore, structural design considerations such as linings, outer layers, and inner layers of protective sheets should be considered for effective shielding in medical institutions. Full article
Show Figures

Figure 1

Figure 1
<p>Shielding design based on radiation intensity (<span class="html-italic">I</span><sub>0</sub>: incident intensity, <span class="html-italic">I</span>: transmitted intensity, <span class="html-italic">μ</span>: absorption coefficient, <span class="html-italic">χ</span>: thickness).</p>
Full article ">Figure 2
<p>Shielding sheets of (<b>A</b>) barium sulfate, (<b>B</b>) bismuth oxide, and (<b>C</b>) tungsten oxide.</p>
Full article ">Figure 3
<p>Evaluation of radiation-shielding performance of shielding sheets.</p>
Full article ">Figure 4
<p>Shielding sheets manufactured using: (<b>A</b>) tungsten oxide, (<b>B</b>) bismuth oxide, (<b>C</b>) barium sulfate, and (<b>D</b>) mixed sheet.</p>
Full article ">
18 pages, 5210 KiB  
Article
Isolation and Characterization of Novel Cellulose Micro/Nanofibers from Lygeum spartum Through a Chemo-Mechanical Process
by Sabrina Ahmima, Nacira Naar, Patryk Jędrzejczak, Izabela Klapiszewska, Łukasz Klapiszewski and Teofil Jesionowski
Polymers 2024, 16(21), 3001; https://doi.org/10.3390/polym16213001 - 25 Oct 2024
Viewed by 757
Abstract
Recent studies have focused on the development of bio-based products from sustainable resources using green extraction approaches, especially nanocellulose, an emerging nanoparticle with impressive properties and multiple applications. Despite the various sources of cellulose nanofibers, the search for alternative resources that replace wood, [...] Read more.
Recent studies have focused on the development of bio-based products from sustainable resources using green extraction approaches, especially nanocellulose, an emerging nanoparticle with impressive properties and multiple applications. Despite the various sources of cellulose nanofibers, the search for alternative resources that replace wood, such as Lygeum spartum, a fast-growing Mediterranean plant, is crucial. It has not been previously investigated as a potential source of nanocellulose. This study investigates the extraction of novel cellulose micro/nanofibers from Lygeum spartum using a two-step method, including both alkali and mechanical treatment as post-treatment with ultrasound, as well as homogenization using water and dilute alkali solution as a solvent. To determine the structural properties of CNFs, a series of characterization techniques was applied. A significant correlation was observed between the Fourier transform infrared (FTIR) spectroscopy and X-ray diffraction (XRD) results. The FTIR results revealed the elimination of amorphous regions and an increase in the energy of the H-bonding modes, while the XRD results showed that the crystal structure of micro/nanofibers was preserved during the process. In addition, they indicated an increase in the crystallinity index obtained with both methods (deconvolution and Segal). Thermal analysis based on thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) confirmed improvement in the thermal properties of the isolated micro/nanofibers. The temperatures of maximum degradation were 335 °C and 347 °C. Morphological analysis using a scanning electron microscope (SEM) and atomic force microscope (AFM) showed the formation of fibers along the axis, with rough and porous surfaces. The findings indicate the potential of Lygeum spartum as a source for producing high-quality micro/nanofibers. A future direction of study is to use the cellulose micro/nanofibers as additives in recycled paper and to evaluate the mechanical properties of the paper sheets, as well as investigate their use in smart paper. Full article
(This article belongs to the Section Biobased and Biodegradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Steps in the preparation of cellulosic micro/nanofibers.</p>
Full article ">Figure 2
<p>FTIR spectra of cellulose-LS, CNF-1, and CNF-2.</p>
Full article ">Figure 3
<p>FTIR results for deconvoluted hydrogen bonding of cellulose-LS (<b>A</b>), CNF-1 (<b>B</b>), and CNF-2 (<b>C</b>) (3800–3000 cm<sup>−1</sup> region). (I) O2–H⋯⋯O6, (II) O3–H⋯⋯O5, (III) O6–H⋯⋯O3.</p>
Full article ">Figure 4
<p>XRD patterns of (<b>A</b>) the isolated micro/nanofibers and (<b>B</b>,<b>C</b>) deconvolutions of micro/nanofibers.</p>
Full article ">Figure 5
<p>(<b>A</b>) TGA, (<b>B</b>) DTG, and (<b>C</b>) DTA curves for cellulose-LS, CNFs-1, and CNFs-2.</p>
Full article ">Figure 6
<p>DSC thermogram of micro/nanofibers.</p>
Full article ">Figure 7
<p>SEM images of (<b>A</b>,<b>A’</b>) cellulose-LS, (<b>B</b>,<b>B’</b>) CNFs-1, and (<b>C</b>,<b>C’</b>) CNFs-2.</p>
Full article ">Figure 8
<p>SEM image of the CNF-2 sample with indications of micro- and nanofibers.</p>
Full article ">Figure 9
<p>AFM images (2D, 3D) of CNFs: (<b>A</b>,<b>A’</b>) CNF-1 and (<b>B</b>,<b>B’</b>) CNF-2.</p>
Full article ">
15 pages, 5545 KiB  
Article
Electroless Copper Patterning on TiO2-Functionalized Mica for Flexible Electronics
by Bozhidar I. Stefanov, Boriana R. Tzaneva, Valentin M. Mateev and Ivo T. Iliev
Appl. Sci. 2024, 14(21), 9780; https://doi.org/10.3390/app14219780 - 25 Oct 2024
Viewed by 753
Abstract
The formation of conductive copper patterns on mica holds promise for developing cost-effective flexible electronics and sensing devices, though it is challenging due to the low adhesion of mica’s atomically flat surface. Herein, we present a wet-chemical method for copper patterning on flexible [...] Read more.
The formation of conductive copper patterns on mica holds promise for developing cost-effective flexible electronics and sensing devices, though it is challenging due to the low adhesion of mica’s atomically flat surface. Herein, we present a wet-chemical method for copper patterning on flexible mica substrates via electroless copper deposition (Cu-ELD). The process involves pre-functionalizing 50 µm thick muscovite mica with a titanium dioxide (TiO2) layer, via a sol–gel dip-coating method with a titanium acetylacetonate-based sol. Photolithography is employed to selectively activate the TiO2-coated mica substrates for Cu-ELD, utilizing in situ photodeposited silver (Ag) nanoclusters as a catalyst. Copper is subsequently plated using a formaldehyde-based Cu-ELD bath, with the duration of deposition primarily determining the thickness and electrical properties of the copper layer. Conductive Cu layers with thicknesses in the 70–130 nm range were formed within 1–2 min of deposition, exhibiting an inverse relationship between plating time and sheet resistance, which ranged from 600 to 300 mΩ/sq. The electrochemical thickening of these layers to 1 μm further reduced the sheet resistance to 27 mΩ/sq. Finally, the potential of Cu-ELD patterning on TiO2-functionalized mica for creating functional sensing devices was demonstrated by fabricating a functional resistance temperature detector (RTD) on the titania surface. Full article
(This article belongs to the Section Chemical and Molecular Sciences)
Show Figures

Figure 1

Figure 1
<p>Schematic of the TiO<sub>2</sub>-functionalization, Ag photodeposition activation patterning, and Cu electroless deposition procedure.</p>
Full article ">Figure 2
<p>Images presenting: (<b>a</b>) schematic of the RTD serpentine topology; (<b>b</b>) photographic image of the resulting Cu-ELD fabricated RTD device.</p>
Full article ">Figure 3
<p>Schematic representation of the temperature-controlled heater setup used in RTD-property determination experiments.</p>
Full article ">Figure 4
<p>Characterization of the Ag/TiO<sub>2</sub>/mica substrates used for Cu-ELD: (<b>a</b>) A representative XRD pattern showing diffraction peaks from the mica substrate; (<b>b</b>) Raman spectra; and (<b>c</b>) UV/Vis transmittance spectra of the pristine, TiO<sub>2</sub>-functionalized, and Ag-activated substrates (photodeposition D<sup>UV</sup> of 2.5 J cm<sup>−2</sup> in all cases).</p>
Full article ">Figure 5
<p>AFM images of the substrates, at a 10 × 10 μm field of view, at different stages of surface treatment: (<b>a</b>) pristine mica surface; (<b>b</b>) TiO<sub>2</sub>-functionalized mica surface; (<b>c</b>) TiO<sub>2</sub>/Mica surface after photodeposition of Ag at D<sup>UV</sup> of 2.5 J cm<sup>−2</sup>.</p>
Full article ">Figure 6
<p>SEM images and EDX spectra of the samples: (<b>a</b>,<b>c</b>) TiO<sub>2</sub>/Mica substrate; (<b>b</b>,<b>d</b>) Cu-ELD layer after 2 min on Ag/TiO<sub>2</sub>/Mica substrate.</p>
Full article ">Figure 6 Cont.
<p>SEM images and EDX spectra of the samples: (<b>a</b>,<b>c</b>) TiO<sub>2</sub>/Mica substrate; (<b>b</b>,<b>d</b>) Cu-ELD layer after 2 min on Ag/TiO<sub>2</sub>/Mica substrate.</p>
Full article ">Figure 7
<p>AFM topography images, at a 10 × 10 μm field of view, for the Cu-ELD layer formed after: (<b>a</b>) 1 min and (<b>b</b>) 2 min of Cu-ELD plating. Both layers were deposited on Ag/TiO<sub>2</sub>/Mica substrates activated with D<sup>UV</sup> of 2.5 J cm<sup>−2</sup>.</p>
Full article ">Figure 8
<p>XRD data showing the Cu (111) reflection, consistent with metallic Cu, for Cu-ELD layers grown on the Ag/TiO<sub>2</sub>/mica substrate after 1, 1.5, and 2 min of deposition.</p>
Full article ">Figure 9
<p>AFM topography images at (<b>a</b>–<b>c</b>) a 100 × 100 μm field of view, and corresponding cross-section plots (<b>d</b>–<b>f</b>) for determining Cu-ELD layer thickness after (<b>a</b>,<b>d</b>) 1 min, (<b>b</b>,<b>e</b>) 1.5 min, and (<b>c</b>,<b>f</b>) 2 min in the Cu-ELD bath. The plots, displayed in panes (<b>d</b>–<b>f</b>), are obtained according to the position and scan direction indicated via yellow arrows in panes (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 10
<p>SEM-EDX analysis of a Cu layer (2 min Cu-ELD), thickened to 1 μm via electrochemical Cu plating: (<b>a</b>) SEM surface morphology; (<b>b</b>) EDX spectrum.</p>
Full article ">Figure 11
<p>Representative data showing the performance of Cu-ELD patterns as resistance temperature detectors for a Cu-ELD serpentine pattern developed on a TiO<sub>2</sub>/Mica substrate during 2 min of electroless deposition: (<b>a</b>) raw data showing the correlation between temperature and direct current resistance; (<b>b</b>) processed data showing the functional dependence used to estimate the temperature coefficient of resistance (TCR).</p>
Full article ">Figure 12
<p>Summary of the effects of Cu-ELD plating time on the RTD functionality of Cu patterns: (<b>a</b>) initial resistance at 30 °C (R30) as a function of deposition time; (<b>b</b>) absolute change in resistance from 30 to 65 °C and TCR as a function of R30.</p>
Full article ">
13 pages, 9014 KiB  
Article
Influence of Synthesis Parameters on Structure and Characteristics of the Graphene Grown Using PECVD on Sapphire Substrate
by Šarūnas Jankauskas, Šarūnas Meškinis, Nerija Žurauskienė and Asta Guobienė
Nanomaterials 2024, 14(20), 1635; https://doi.org/10.3390/nano14201635 - 12 Oct 2024
Viewed by 934
Abstract
The high surface area and transfer-less growth of graphene on dielectric materials is still a challenge in the production of novel sensing devices. We demonstrate a novel approach to graphene synthesis on a C-plane sapphire substrate, involving the microwave plasma-enhanced chemical vapor deposition [...] Read more.
The high surface area and transfer-less growth of graphene on dielectric materials is still a challenge in the production of novel sensing devices. We demonstrate a novel approach to graphene synthesis on a C-plane sapphire substrate, involving the microwave plasma-enhanced chemical vapor deposition (MW-PECVD) technique. The decomposition of methane, which is used as a precursor gas, is achieved without the need for remote plasma. Raman spectroscopy, atomic force microscopy and resistance characteristic measurements were performed to investigate the potential of graphene for use in sensing applications. We show that the thickness and quality of graphene film greatly depend on the CH4/H2 flow ratio, as well as on chamber pressure during the synthesis. By varying these parameters, the intensity ratio of Raman D and G bands of graphene varied between ~1 and ~4, while the 2D to G band intensity ratio was found to be 0.05–0.5. Boundary defects are the most prominent defect type in PECVD graphene, giving it a grainy texture. Despite this, the samples exhibited sheet resistance values as low as 1.87 kΩ/□. This reveals great potential for PECVD methods and could contribute toward efficient and straightforward graphene growth on various substrates. Full article
(This article belongs to the Section Nanofabrication and Nanomanufacturing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the PECVD chamber during graphene synthesis on C-sapphire substrate, (<b>b</b>) illustration of C-AFM analysis, (<b>c</b>) illustration of four-point probe analysis for sheet resistance measurement and (<b>d</b>) schematic representation of PECVD graphene synthesis stages. Atom spacing, electrode placement and dimensions are not to scale for clarity.</p>
Full article ">Figure 2
<p>(<b>a</b>) Comparison of Raman spectra of PECVD graphene on C−plane sapphire (violet) and SiO<sub>2</sub> (orange), (<b>b</b>) colormap of I<sub>2D</sub>/I<sub>G</sub> ratio vs. synthesis parameters, (<b>c</b>) I<sub>D</sub>/I<sub>G</sub> ratio vs. I<sub>2D</sub>/I<sub>G</sub> ratio, (<b>d</b>) colormap of I<sub>D</sub>/I<sub>G</sub> ratio vs. synthesis parameters. The samples were differentiated by shapes and color, where light green represents samples that belong to the F set and light blue represents samples that belong to the P set. Sample F3P2, which belongs to both sets, is colored red. Additional samples were produced (set S) for colormap space expansion.</p>
Full article ">Figure 3
<p>(<b>a</b>) Pos<sub>2D</sub> vs. gas-flow ratio plot, (<b>b</b>) Pos<sub>2D</sub> vs. pressure plot, (<b>c</b>) FWHM<sub>2D</sub> vs. gas-flow ratio plot, (<b>d</b>) FWHM<sub>2D</sub> vs. pressure plot of F and P samples. Conveniently, samples were given the same shapes and colors for distinction.</p>
Full article ">Figure 4
<p>Pos<sub>2D</sub> vs. Pos<sub>G</sub> plot showing vector decomposition analysis of F and P sample sets. Conveniently, samples were given the same shapes and colors for distinction.</p>
Full article ">Figure 5
<p>(<b>a</b>) Morphology of graphene synthesized using a CH<sub>4</sub>/H<sub>2</sub> flow ratio of 35/65 and a pressure of 10 (R<sub>q</sub> indicates the root mean square value of surface roughness), (<b>b</b>) R<sub>q</sub> vs. gas-flow ratio plot, (<b>c</b>) R<sub>q</sub> vs. pressure plot. The colors and shapes follow the same sample pattern.</p>
Full article ">Figure 6
<p>(<b>a</b>) Two-dimensional (2D) and G band ratio with respect to surface roughness, (<b>b</b>) plot showing correlation with surface roughness. The colors and shapes follow the same sample pattern.</p>
Full article ">Figure 7
<p>Conductive atomic force microscopy maps of four selected samples with varying synthesis conditions. (<b>a</b>) F2 sample, (<b>b</b>) F5 sample, (<b>c</b>) P1 sample, (<b>d</b>) P5 sample. I<sub>q</sub> represents the root mean square of surface conductivity values.</p>
Full article ">Figure 8
<p>(<b>a</b>) Resistance characteristics vs. I<sub>2D</sub>/I<sub>G</sub> plot showing conductance and sheet resistance variation with different thicknesses of graphene, (<b>b</b>) resistance characteristics vs. I<sub>D</sub>/I<sub>G</sub> plot showing conductance and sheet resistance variation based on defects.</p>
Full article ">
19 pages, 1983 KiB  
Article
Numerical Analysis of Liquid Hydrogen Atomization in a Premixing Tube Using a Volume of Fluid-to-Discrete Particle Model Approach
by Abdalazeem Adam, Weifeng He, Dong Han, Yuxin Fan and Omer Musa
Aerospace 2024, 11(10), 832; https://doi.org/10.3390/aerospace11100832 - 10 Oct 2024
Viewed by 791
Abstract
This paper examines the atomization characteristics of liquid hydrogen fuel in a premixing tube under different operating conditions. Hydrogen fuel’s unique injection morphology and atomization behavior were analyzed using the Volume of Fluid-to-Discrete Particle Model (VOF to DPM) approach, coupled with the SST [...] Read more.
This paper examines the atomization characteristics of liquid hydrogen fuel in a premixing tube under different operating conditions. Hydrogen fuel’s unique injection morphology and atomization behavior were analyzed using the Volume of Fluid-to-Discrete Particle Model (VOF to DPM) approach, coupled with the SST kω turbulence model and adaptive mesh refinement. The study revealed that the breakup and transformation of liquid surfaces into particles are significantly impacted by varying air velocities and injection pressure. Specifically, higher air velocities caused the liquid sheet to lengthen and narrow due to intensified vortices. However, the breakup was delayed at higher velocities, occurring at distances of 0.037 m and 0.043 m for air velocities of 10 m/s and 20 m/s, respectively. The research also highlights the significant role that injection pressure plays in fluid sheet breakup. Higher pressures promote better atomization and fuel–lair mixing, resulting in more particles with increased diameters. Notably, the fluid sheet exhibited a small angle of about 43.79° when using the velocity component corresponding to p1 = 0.5 MPa. Similarly, for p2 = 1 MPa and p3 = 2 MPa, the angles were measured to be approximately 47.5° and 49.5°, respectively. Additionally, the study observed that the injection expands in length and diameter as time progresses, indicating fuel dispersion. These insights have significant implications for the design principles of injectors in power generation technologies that utilize liquid hydrogen fuel. Full article
(This article belongs to the Section Aeronautics)
Show Figures

Figure 1

Figure 1
<p>Computational domain of the premixer tube.</p>
Full article ">Figure 2
<p>Schematic diagram of structured grids of the swirl premixer tube.</p>
Full article ">Figure 3
<p>Grid independence study using velocity magnitude at x = 0.08 m with four types of grids.</p>
Full article ">Figure 4
<p>A comparison of injected fuel shapes: (<b>a</b>) reference [<a href="#B20-aerospace-11-00832" class="html-bibr">20</a>]; (<b>b</b>) the injection of this study.</p>
Full article ">Figure 5
<p>Iso-surfaces represent the fluid sheet formation process at nine different time intervals from 0.15 ms to 2.4 ms.</p>
Full article ">Figure 6
<p>Iso-surfaces represent the fluid sheet breakup process at nine different time intervals from 0.15 ms to 3.15 ms and particles with different diameters.</p>
Full article ">Figure 7
<p>Iso-surfaces of hydrogen fluid sheet at different air inlet velocities and times (1 and 3 ms): (<b>a</b>,<b>b</b>) air inlet velocity of 0 m/s; (<b>c</b>,<b>d</b>) air inlet velocity of 10 m/s; and (<b>e</b>,<b>f</b>) air inlet velocity of 20 m/s.</p>
Full article ">Figure 8
<p>Contour of velocity at t = 2 ms in the center of the premixing tube: (<b>a</b>) air inlet velocity of 0 m/s; (<b>b</b>) air inlet velocity of 10 m/s; and (<b>c</b>) air inlet velocity of 20 m/s.</p>
Full article ">Figure 9
<p>Diameter and locations of particles at different air inlet velocities and times (1 and 3 ms): (<b>a</b>,<b>b</b>) air inlet velocity of 0 m/s; (<b>c</b>,<b>d</b>) air inlet velocity of 10 m/s; and (<b>e</b>,<b>f</b>) air inlet velocity of 20 m/s.</p>
Full article ">Figure 10
<p>Effect of various air inlet velocities along the flow direction: (<b>a</b>) the velocity magnitude and (<b>b</b>) turbulent kinetic energy.</p>
Full article ">Figure 11
<p>Iso-surface of the fluid sheet for the different swirl injection radial, tangential, and axial velocity components: (<b>P1</b>) 2.58 m/s, 9.89 m/s, and 18.93 m/s; (<b>P2</b>) 5.1 m/s, 36 m/s, and 42.2 m/s; (<b>P3</b>) 8 m/s, 48 m/s, and 42 m/s.</p>
Full article ">Figure 12
<p>Particles at the same flow time for the different swirl injection radial, tangential, and axial velocity components: (<b>a</b>) front view of the tube and (<b>b</b>) side view of the tube.</p>
Full article ">Figure 13
<p>Iso-surface of fluid sheet and ligament breakup for the two different swirl injection radial, tangential, and axial velocity components at the same flow time: (<b>a</b>) 2.58 m/s, 9.89 m/s, and 18.93 m/s; (<b>b</b>) 4.12 m/s, 17.06 m/s, and 29.96 m/s.</p>
Full article ">Figure 14
<p>Particle number for the two different swirl injection radial, tangential, and axial velocity components at the same flow time: (<b>a</b>) 2.58 m/s, 9.89 m/s, and 18.93 m/s; (<b>b</b>) 4.12 m/s, 17.06 m/s, and 29.96 m/s.</p>
Full article ">Figure 15
<p>Contour of velocity highlighting the differences between the two cases in the center of the premixing tube: (<b>a</b>) 2.58 m/s, 9.89 m/s, and 18.93 m/s; (<b>b</b>) 4.12 m/s, 17.06 m/s, and 29.96 m/s.</p>
Full article ">Figure 16
<p>(<b>a</b>) Velocity magnitude and (<b>b</b>) turbulent kinetic energy along a vertical line at a distance of 9 mm away from the fuel inlet.</p>
Full article ">
8 pages, 1342 KiB  
Article
Metal-Cation-Induced Tiny Ripple on Graphene
by Yingying Huang, Hanlin Li, Liuyuan Zhu, Yongshun Song and Haiping Fang
Nanomaterials 2024, 14(19), 1593; https://doi.org/10.3390/nano14191593 - 2 Oct 2024
Viewed by 820
Abstract
Ripples on graphene play a crucial role in manipulating its physical and chemical properties. However, producing ripples, especially at the nanoscale, remains challenging with current experimental methods. In this study, we report that tiny ripples in graphene can be generated by the adsorption [...] Read more.
Ripples on graphene play a crucial role in manipulating its physical and chemical properties. However, producing ripples, especially at the nanoscale, remains challenging with current experimental methods. In this study, we report that tiny ripples in graphene can be generated by the adsorption of a single metal cation (Na+, K+, Mg2+, Ca2+, Cu2+, Fe3+) onto a graphene sheet, based on the density functional theory calculations. We attribute this to the cation–π interaction between the metal cation and the aromatic rings on the graphene surface, which makes the carbon atoms closer to metal ions, causing deformation of the graphene sheet, especially in the out-of-plane direction, thereby creating ripples. The equivalent pressures applied to graphene sheets in out-of-plane direction, generated by metal cation–π interactions, reach magnitudes on the order of gigapascals (GPa). More importantly, the electronic and mechanical properties of graphene sheets are modified by the adsorption of various metal cations, resulting in opened bandgaps and enhanced rigidity characterized by a higher elastic modulus. These findings show great potential for applications for producing ripples at the nanoscale in graphene through the regulation of metal cation adsorption. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Structural configuration of a metal cation adsorbed on a graphene sheet. M<sup>n+</sup> denotes the metal cation, including Na<sup>+</sup>, K<sup>+</sup>, Mg<sup>2+</sup>, Ca<sup>2+</sup>, Cu<sup>2+</sup>, and Fe<sup>3+</sup>. (<b>b</b>) Distances between various metal cations and the corresponding graphene sheets. (<b>c</b>) Adsorption energies of the various metal cations on the graphene sheet. (<b>d</b>) Number of electrons transferred from the graphene sheet to the metal cations.</p>
Full article ">Figure 2
<p>Partial electron density of states near the Fermi level of (<b>a</b>) Cu<sup>2+</sup>@graphene and (<b>b</b>) Fe<sup>3+</sup>@graphene.</p>
Full article ">Figure 3
<p>(<b>a</b>) Out-of-plane deformation (Δ<span class="html-italic">Z</span>) of rippled graphene induced by the adsorption of various metal cations. (<b>b</b>) Equivalent pressure (<span class="html-italic">P</span>) exerted by the metal cation on the graphene sheet.</p>
Full article ">Figure 4
<p>(<b>a</b>) Band structure and density of states (DOS) for Fe<sup>3+</sup> adsorbed on the graphene sheet. <span class="html-italic">E</span><sub>g</sub> represents the band gap. (<b>b</b>) Band gaps of rippled graphene sheets induced by various metal cations.</p>
Full article ">
12 pages, 2000 KiB  
Article
Adsorption of Atomic Hydrogen on Hydrogen Boride Sheets Studied by Photoelectron Spectroscopy
by Heming Yin, Jingmin Tang, Kazuki Yamaguchi, Haruto Sakurai, Yuki Tsujikawa, Masafumi Horio, Takahiro Kondo and Iwao Matsuda
Materials 2024, 17(19), 4806; https://doi.org/10.3390/ma17194806 - 29 Sep 2024
Viewed by 1116
Abstract
Hydrogen boride (HB) sheets are emerging as a promising two-dimensional (2D) boron material, with potential applications as unique electrodes, substrates, and hydrogen storage materials. The 2D layered structure of HB was successfully synthesized using an ion-exchange method. The chemical bonding and structure of [...] Read more.
Hydrogen boride (HB) sheets are emerging as a promising two-dimensional (2D) boron material, with potential applications as unique electrodes, substrates, and hydrogen storage materials. The 2D layered structure of HB was successfully synthesized using an ion-exchange method. The chemical bonding and structure of the HB sheets were investigated using Fourier Transform Infrared (FT–IR) spectroscopy and Transmission Electron Microscopy (TEM), respectively. X-ray photoelectron spectroscopy (XPS) was employed to study the chemical states and transformation of the components before and after atomic hydrogen adsorption, thereby elucidating the atomic hydrogen adsorption process on HB sheets. Our results indicate that, upon atomic hydrogen adsorption onto the HB sheets, the B-H-B bonds were broken and converted into B-H bonds. This research highlights and demonstrates the changes in chemical states and component transformations of the boron element on the HB sheets’ surface before and after atomic hydrogen adsorption, thus providing a clearer understanding of the unique bonding and structural characteristics of the HB sheets. Full article
(This article belongs to the Special Issue Development of Boron-Based Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Schematic synthesizing process of HB sheets by ion-exchange. (<b>b</b>) FT–IR spectrum of HB sheets. (<b>c</b>) TEM images of HB sheets; inset shows the region enclosed by the red square. A winding structure is traced by a red curve. (<b>d</b>) Structure model of HB sheets.</p>
Full article ">Figure 2
<p>XPS B 1s core-level spectra for the HB sheets with photon energy of 285 eV. (<b>a</b>) Background subtraction of XPS B 1s spectra: 3000 L atomic hydrogen adsorbed on HB sheets (<b>above</b>) and initial HB sheets (<b>below</b>). (<b>b</b>) Curve fitting of the B 1s XPS spectra.</p>
Full article ">Figure 3
<p>XPS B 1s core-level spectra for the HB sheets with photon energy of 630 eV. (<b>a</b>) Background subtraction of XPS B 1s spectra: 3000 L atomic hydrogen adsorbed on HB sheets (<b>above</b>) and initial HB sheets (<b>below</b>). (<b>b</b>) Curve fitting of the B 1s XPS spectra.</p>
Full article ">Figure 4
<p>XPS B 1s core-level spectra for bulk boron with photon energy of 285 eV. (<b>a</b>) Background subtraction of XPS B 1s spectra: 3000 L atomic hydrogen adsorbed on HB sheets (<b>above</b>) and initial HB sheets (<b>below</b>). (<b>b</b>) Curve fitting of the B 1s XPS spectra.</p>
Full article ">Figure 5
<p>XPS B 1s core-level spectra for bulk boron with photon energy of 630 eV. (<b>a</b>) Background subtraction of XPS B 1s spectra: 3000 L atomic hydrogen adsorbed on HB sheets (<b>above</b>) and initial HB sheets (<b>below</b>). (<b>b</b>) Curve fitting of the B 1s XPS spectra.</p>
Full article ">
11 pages, 3858 KiB  
Article
Mica Lattice Orientation of Epitaxially Grown Amyloid β25–35 Fibrils
by György G. Ferenczy, Ünige Murvai, Lívia Fülöp and Miklós Kellermayer
Int. J. Mol. Sci. 2024, 25(19), 10460; https://doi.org/10.3390/ijms251910460 - 28 Sep 2024
Viewed by 726
Abstract
β-amyloid (Aβ) peptides form self-organizing fibrils in Alzheimer’s disease. The biologically active, toxic Aβ25–35 fragment of the full-length Aβ-peptide forms a stable, oriented filament network on the mica surface with an epitaxial mechanism at the timescale of seconds. While many of the structural [...] Read more.
β-amyloid (Aβ) peptides form self-organizing fibrils in Alzheimer’s disease. The biologically active, toxic Aβ25–35 fragment of the full-length Aβ-peptide forms a stable, oriented filament network on the mica surface with an epitaxial mechanism at the timescale of seconds. While many of the structural and dynamic features of the oriented Aβ25–35 fibrils have been investigated before, the β-strand arrangement of the fibrils and their exact orientation with respect to the mica lattice remained unknown. By using high-resolution atomic force microscopy, here, we show that the Aβ25–35 fibrils are oriented along the long diagonal of the oxygen hexagon of mica. To test the structure and stability of the oriented fibrils further, we carried out molecular dynamics simulations on model β-sheets. The models included the mica surface and a single fibril motif built from β-strands. We show that a sheet with parallel β-strands binds to the mica surface with its positively charged groups, but the C-terminals of the strands orient upward. In contrast, the model with antiparallel strands preserves its parallel orientation with the surface in the molecular dynamics simulation, suggesting that this model describes the first β-sheet layer of the mica-bound Aβ25–35 fibrils well. These results pave the way toward nanotechnological construction and applications for the designed amyloid peptides. Full article
(This article belongs to the Special Issue The Role of Environment in Amyloid Aggregation: 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Height-contrast AFM image of a mica surface on which Aβ25–35 fibrils were grown by epitaxy. The fibrils are of a more-or-less uniform width and height, straight, and are oriented in three main directions that reflect the underlying hexagonal lattice structure of mica. Color bar indicates the topographical height scale of the AFM image.</p>
Full article ">Figure 2
<p>Schematics of the possible arrangements of the epitaxially grown Aβ25–35 fibrils on the mica surface with respect to the mica lattice. (<b>a</b>) Fibril axis is parallel to the edge of the oxygen-atom hexagon (i.e., oriented along its long diagonal) and hence perpendicular to the edge of the silicon-atom hexagon. (<b>i</b>,<b>ii</b>) Parallel and antiparallel β-strand orientations, respectively. (<b>b</b>) Fibril axis is perpendicular to the edge of the oxygen-atom hexagon (i.e., oriented along its short diagonal) and hence parallel with the edge of the silicon-atom hexagon. (<b>i</b>,<b>ii</b>) Parallel and antiparallel β-strand orientations, respectively. Altogether, the structural scenarios described in the text are as follows: (<b>ai</b>) scenario 1, (<b>aii</b>) scenario 2, (<b>bi</b>) scenario 3, and (<b>bii</b>) scenario 4.</p>
Full article ">Figure 3
<p>AFM analysis of the Aβ25–35 fibril orientation with respect to the mica crystal lattice. (<b>a</b>) Overview height-contrast AFM image of Aβ25–35 fibrils. (<b>b</b>) Magnified AFM image of the area boxed in (<b>a</b>). (<b>c</b>) Atomic-resolution AFM image of a 10-by-10 nm area in the sample. (<b>d</b>) Overlay of images (<b>b</b>,<b>c</b>), for a better comparison of the fibril and lattice orientations. (<b>e</b>) Schematics of the hexagonal mica structure with the overlain Aβ25–35 fibril orientations. The lattice orientation in the schematic diagram is essentially identical to that in the AFM image. Accordingly, the fibril axis is parallel to the oxygen-atom hexagon and hence perpendicular to the edge of the silicon-atom hexagon. Fibril orientations are shown by green arrows and a selected orientation is indicated by a row of more darkly shaded O atoms. Color bars indicate the topographical height scales of the AFM images.</p>
Full article ">Figure 4
<p>Molecular dynamics simulation of a twelve-strand-long, single-layer Aβ25–35 β-sheet on mica. (<b>a</b>,<b>b</b>) Initial and final arrangements of an antiparallel β-sheet after a 100 ns long simulation, respectively. (<b>c</b>,<b>d</b>) Initial and final arrangements of a parallel β-sheet after a 100 ns long simulation, respectively. Mica is shown here with a space-filling model: Si coral, Al mustard, O red, K pink. Peptide strands are shown with ribbons and in thin tube representation. Backbone hydrogen bonds are indicated with yellow dotted lines.</p>
Full article ">Figure 5
<p>Summary of the most plausible structural arrangement of the epitaxially grown Aβ25–35 fibril on mica. (<b>a</b>) Hexagonal arrangement of the oxygen and silicon atoms in the surface crystal lattice of mica. (<b>b</b>) Section of an Aβ25–35 fibril. The fibril axis orientation is indicated with a yellow dotted arrow. Selected O-hexagons are indicated by black dashed lines and selected Si-hexagons by black solid lines. (<b>c</b>) Perspective view of the first β-sheet of the Aβ25–35 fibril, in which the β-strands are oriented antiparallel.</p>
Full article ">
17 pages, 6501 KiB  
Article
Enhancing Mechanical Properties of Graphene/Aluminum Nanocomposites via Microstructure Design Using Molecular Dynamics Simulations
by Zhonglei Ma, Hongding Wang, Yanlong Zhao, Zhengning Li, Hong Liu, Yizhao Yang and Zigeng Zhao
Materials 2024, 17(18), 4552; https://doi.org/10.3390/ma17184552 - 16 Sep 2024
Viewed by 1208
Abstract
This study explores the mechanical properties of graphene/aluminum (Gr/Al) nanocomposites through nanoindentation testing performed via molecular dynamics simulations in a large-scale atomic/molecular massively parallel simulator (LAMMPS). The simulation model was initially subjected to energy minimization at 300 K, followed by relaxation for 50 [...] Read more.
This study explores the mechanical properties of graphene/aluminum (Gr/Al) nanocomposites through nanoindentation testing performed via molecular dynamics simulations in a large-scale atomic/molecular massively parallel simulator (LAMMPS). The simulation model was initially subjected to energy minimization at 300 K, followed by relaxation for 50 ps under the NPT ensemble, wherein the number of atoms (N), simulation temperature (T), and pressure (P) were conserved. After the model was fully relaxed, loading and unloading simulations were performed. This study focused on the effects of the Gr arrangement with a brick-and-mortar structure and incorporation of high-entropy alloy (HEA) coatings on mechanical properties. The findings revealed that Gr sheets (GSs) significantly impeded dislocation propagation, preventing the dislocation network from penetrating the Gr layer within the plastic zone. However, interactions between dislocations and GSs in the Gr/Al nanocomposites resulted in reduced hardness compared with that of pure aluminum. After modifying the arrangement of GSs and introducing HEA (FeNiCrCoAl) coatings, the elastic modulus and hardness of the Gr/Al nanocomposites were 83 and 9.5 GPa, respectively, representing increases of 21.5% and 17.3% compared with those of pure aluminum. This study demonstrates that vertically oriented GSs in combination with HEA coatings at a mass fraction of 3.4% significantly enhance the mechanical properties of the Gr/Al nanocomposites. Full article
(This article belongs to the Section Materials Simulation and Design)
Show Figures

Figure 1

Figure 1
<p>Schematic of the nanoindentation models: pure aluminum (<b>a</b>), Gr/Al-level3 (<b>b</b>), Gr/Al-vertical3 (<b>c</b>), HEA/Gr/Al-level3 (<b>d</b>), and HEA/Gr/Al-vertical3 (<b>e</b>).</p>
Full article ">Figure 2
<p>Dislocation distributions observed along the x-axis for the five models, namely pure aluminum, Gr/Al-level3, Gr/Al-vertical3, HEA/Gr/Al-level3, and HEA/Gr/Al-vertical3, at indenter displacements of d = 15, 20, 25, 30, and 35 Å. Dislocations are colored according to their Burgers vector. Green: 1/6&lt;112&gt;; Dark blue: 1/2&lt;110&gt;; Pink: 1/6&lt;110&gt;; Yellow: 1/3&lt;100&gt;; Bright blue: 1/3&lt;111&gt;; Red: others.</p>
Full article ">Figure 3
<p>Side views of the y–z-planes of Gr/Al-level3 (<b>a</b>) and HEA/Gr/Al-level3 (<b>b</b>) at indentation depths of 0, 20, 25, and 30 Å. The viewing direction is along the x-axis.</p>
Full article ">Figure 4
<p>Side views of the y–z-planes of Gr/Al-vertical3 (<b>a</b>) at indentation depths of 0, 26, 27, 28, and 29 Å and HEA/Gr/Al-vertical3 (<b>b</b>) at indentation depths of 0, 9, 11, 13, and 16 Å. The viewing direction is along the x-axis.</p>
Full article ">Figure 5
<p>In-plane height profiles of Gr at an indentation depth of 30 Å for models Gr/Al-level3 (<b>a</b>), HEA/Gr/Al-level3 (<b>b</b>), Gr/Al-vertical3 (<b>c</b>), and HEA/Gr/Al-vertical3 (<b>d</b>).</p>
Full article ">Figure 6
<p>Evolution of total dislocation length (<b>a</b>) and indentation force (<b>b</b>) with indenter displacement.</p>
Full article ">Figure 7
<p>Hardness values vs. indentation depths.</p>
Full article ">Figure 8
<p>Force–displacement curves at the unloading stage.</p>
Full article ">Figure 9
<p>Distribution of dislocation lines and defect atoms at different indenter displacements during the unloading stage. Dislocations are colored according to their Burgers vector. Green: 1/6&lt;112&gt;; Dark blue: 1/2&lt;110&gt;; Pink: 1/6&lt;110&gt;; Yellow: 1/3&lt;100&gt;; Bright blue: 1/3&lt;111&gt;; Red: others. (<b>a</b>) pure aluminum; (<b>b</b>) Gr/Al-level3; (<b>c</b>) Gr/Al-vertical3; (<b>d</b>) HEA/Gr/Al-level3; (<b>e</b>) HEA/Gr/Al-vertical3.</p>
Full article ">Figure 10
<p>Reduced Young’s modulus of the five models.</p>
Full article ">Figure 11
<p>Dislocation length–indenter displacement curves. (<b>a</b>) pure aluminum; (<b>b</b>) Gr/Al-vertical3; (<b>c</b>) HEA/Gr/Al-level3; (<b>d</b>) HEA/Gr/Al-vertical3.</p>
Full article ">
20 pages, 4982 KiB  
Article
Effects of Soy Protein Isolate and Inulin Conjugate on Gel Properties and Molecular Conformation of Spanish Mackerel Myofibrillar Protein
by Wei Wang, Sirui Ma, Qing Shao and Shumin Yi
Foods 2024, 13(18), 2920; https://doi.org/10.3390/foods13182920 - 15 Sep 2024
Viewed by 958
Abstract
The gel properties and molecular conformation of Spanish mackerel myofibrillar protein (MP) induced by soy protein isolate–inulin conjugates (SPI–inulin conjugates) were investigated. The addition of SPI–inulin conjugates significantly enhanced the quality of the protein gel. An analysis of different additives was conducted to [...] Read more.
The gel properties and molecular conformation of Spanish mackerel myofibrillar protein (MP) induced by soy protein isolate–inulin conjugates (SPI–inulin conjugates) were investigated. The addition of SPI–inulin conjugates significantly enhanced the quality of the protein gel. An analysis of different additives was conducted to assess their impact on the gel strength, texture, water-holding capacity (WHC), water distribution, intermolecular force, dynamic rheology, Raman spectrum, fluorescence spectrum, and microstructure of MP. The results demonstrated a substantial improvement in the strength and water retention of the MP gel with the addition of the conjugate. Compared with the control group (MP), the gel strength increased from 35.18 g·cm to 41.90 g·cm, and WHC increased from 36.80% to 52.67% with the inclusion of SPI–inulin conjugates. The hydrogen bond content was notably higher than that of other groups, and hydrophobic interaction increased from 29.30% to 36.85% with the addition of SPI–inulin conjugates. Furthermore, the addition of the conjugate altered the secondary structure of the myofibrillar gel, with a decrease in α-helix content from 62.91% to 48.42% and an increase in β-sheet content from 13.40% to 24.65%. Additionally, the SPI–inulin conjugates led to a significant reduction in the endogenous fluorescence intensity of MP. Atomic force microscopy (AFM) results revealed a substantial increase in the Rq value from 8.21 nm to 20.21 nm. Adding SPI and inulin in the form of conjugates is an effective method to improve the gel properties of proteins, which provides important guidance for the study of adding conjugates to surimi products. It has potential application prospects in commercial surimi products. Full article
(This article belongs to the Section Food Physics and (Bio)Chemistry)
Show Figures

Figure 1

Figure 1
<p>Effects of different additives on gel strength of myofibrillar gel. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effects of different additives on water retention of myofibrillar gel. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effects of different additives on T<sub>2</sub> relaxation time (<b>A</b>) and the peak area ratio (<b>B</b>) of myofibrillar gel. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Effects of different additives on the chemical force of myofibrillar protein gel. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Effects of different additives on the Raman spectra (<b>A</b>) and protein secondary (<b>B</b>) structure content of myofibrillar protein gel. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Effects of different additives on energy storage modulus (<b>A</b>) and loss modulus (<b>B</b>) of myofibrillar protein. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively.</p>
Full article ">Figure 6 Cont.
<p>Effects of different additives on energy storage modulus (<b>A</b>) and loss modulus (<b>B</b>) of myofibrillar protein. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively.</p>
Full article ">Figure 7
<p>Effects of different additives on the endogenous fluorescence spectra of myofibrillar protein. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively.</p>
Full article ">Figure 8
<p>Effects of different additives on atomic force microscopy of myofibrillar protein. Note: MP: myofibrillar protein; MP-S: myofibrillar protein with 0.6% SPI; MP-I: myofibrillar protein with 0.6% inulin, MP-M: myofibrillar protein with 0.6% mixtures of SPI and inulin; MP-C: myofibrillar protein with 0.6% SPI–inulin conjugates, respectively. Different lowercase letters indicate a significant difference (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 9
<p>Correlation analysis of gel strength, texture, water distribution, water retention, chemical force, microstructure, and secondary structure parameters of myofibrillar protein gel.</p>
Full article ">
14 pages, 19586 KiB  
Article
Advanced Electrospun Composites Based on Polycaprolactone Fibers Loaded with Micronized Tungsten Powders for Radiation Shielding
by Chiara Giuliani, Ilaria De Stefano, Mariateresa Mancuso, Noemi Fiaschini, Luis Alexander Hein, Daniele Mirabile Gattia, Elisa Scatena, Eleonora Zenobi, Costantino Del Gaudio, Federica Galante, Giuseppe Felici and Antonio Rinaldi
Polymers 2024, 16(18), 2590; https://doi.org/10.3390/polym16182590 - 13 Sep 2024
Cited by 1 | Viewed by 1220
Abstract
Exposure to high levels of radiation can cause acute, long-term health effects, such as acute radiation syndrome, cancer, and cardiovascular disease. This is an important occupational hazard in different fields, such as the aerospace and healthcare industry, as well as a crucial burden [...] Read more.
Exposure to high levels of radiation can cause acute, long-term health effects, such as acute radiation syndrome, cancer, and cardiovascular disease. This is an important occupational hazard in different fields, such as the aerospace and healthcare industry, as well as a crucial burden to overcome to boost space applications and exploration. Protective bulky equipment made of heavy metals is not suitable for many advanced purporses, such as mobile devices, wearable shields, and manned spacecrafts. In the latter case, the in-space manufacturing of protective shields is highly desirable and remains an unmet need. Composites made of polymers and high atomic number fillers are potential means for radiation protection due to their low weight, good flexibility, and good processability. In the present work, we developed electrospun composites based on polycaprolactone (polymer matrix) and tungsten powder for application as shielding materials. Electrospinning is a versatile technology that is easily scalable at an industrial level and allows obtaining very lightweight, flexible sheet materials for wearables. By controlling tungsten powder size, we engineered homogeneous, stable and processable suspensions to fabricate radiation composite shielding sheets. The shielding capability was assessed by an in vivo model on prototype composite sheets containing 80 w% of W filler in a polycaprolactone (PCL) fibrous matrix by means of irradiation tests (X-rays) on mice. The obtained results are promising; as expected, the shielding effectivity of the developed composite material increases with the thickness/number of stacked layers. It is worth noting that a thin barrier consisting of 24 layers of the innovative shielding material reduces the extent of apoptosis by 1.5 times compared to the non-shielded mice. Full article
(This article belongs to the Special Issue Advances in Functional Polymers and Composites)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the experimental setup (<b>a</b>), custom-built 3D printed PLA protective device for in vivo test mounted on a 3 mm thick solid lead sheet (<b>b</b>); PLA device wrapped up in 12 shielding sheets of PCL/W (<b>c</b>); PLA device wrapped up in 24 shielding sheets of PCL/W (<b>d</b>); schematic representation of a section of a cerebellum at 2 days of age (<b>e</b>); antero-dorsal cardinal lobe of the cerebellum (<b>f</b>). The dashed black line externally outlines the EGL. Figure in (<b>a</b>,<b>e</b>) was obtained by Biorender.</p>
Full article ">Figure 2
<p>The SEM image of the commercial W powder is as follows: before (<b>a</b>,<b>b</b>), after 16 h (<b>c</b>,<b>d</b>) and 26 h (<b>e</b>,<b>f</b>) of ball milling.</p>
Full article ">Figure 3
<p>Unstable suspension prepared using PCL and the commercial W powder (<b>a</b>), and stable suspension based on PCL and the optimized W powder (<b>b</b>).</p>
Full article ">Figure 4
<p>Micro-fibrous PCL sheets characterized by different W content (w% with respect to the PCL polymer).</p>
Full article ">Figure 5
<p>Microfibrous PCL/W sheets characterized as 10 w% (<b>a</b>), 40 w% (<b>b</b>), and 60 w% (<b>c</b>) of W with respect to the PCL polymer.</p>
Full article ">Figure 6
<p>EDS analysis of the PCL/W electrospun sheets.</p>
Full article ">Figure 7
<p>The apoptotic rate in EGL of irradiated mice under different experimental conditions. (<b>a</b>–<b>d</b>) Representative images of the EGL in the antero-dorsal cardinal lobe region of the cerebellum at postnatal day 2 (P2); Hematoxylin &amp; Eosin staining; 40× magnification. The inset in (<b>a</b>) shows pyknotic nuclei at higher magnification (100×) indicative of apoptosis in WB irradiated mice; (<b>e</b>) a graphical representation of the percentage of apoptotic cells in the different experimental groups. *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 8
<p>(<b>a</b>) Cerebellum at P2 labeled with activated caspase 3; the EGL, where the cells undergoing apoptosis reside, is colored red (2× magnification). (<b>b</b>) Detail of the EGL at higher magnification (40×). (<b>c</b>) Graphical representation of the quantification of the signal related to the antero-dorsal cardinal lobe region. * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
Back to TopTop