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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (230)

Search Parameters:
Keywords = guided atom

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 11855 KiB  
Review
Effective Factors for Optimizing Metallophthalocyanine-Based Optoelectronic Devices: Surface—Molecule Interactions
by Sakineh Akbari Nia, Aleksandra Tomaszowska, Paulina Powroźnik and Maciej Krzywiecki
Molecules 2025, 30(3), 471; https://doi.org/10.3390/molecules30030471 - 22 Jan 2025
Viewed by 342
Abstract
As a promising structure for fabricating inorganic—organic-based optoelectronic devices, metal—metallophthalocyanine (MPc) hybrid layers are highly important to be considered. The efficient charge injection and transport across the metal/MPc interface are strictly dependent on the precise molecular orientation of the MPcs. Therefore, the efficiency [...] Read more.
As a promising structure for fabricating inorganic—organic-based optoelectronic devices, metal—metallophthalocyanine (MPc) hybrid layers are highly important to be considered. The efficient charge injection and transport across the metal/MPc interface are strictly dependent on the precise molecular orientation of the MPcs. Therefore, the efficiency of MPc-based optoelectronic devices strictly depends on the adsorption and orientation of the organic MPc on the inorganic metal substrate. The current review aims to explore the effect of the terminated atoms or surface atoms as an internal stimulus on molecular adsorption and orientation. Here, we investigate the adsorption of five different phthalocyanine molecules—free-based phthalocyanine (H2Pc), copper phthalocyanine (CuPc), iron phthalocyanine (FePc), cobalt phthalocyanine (CoPc), vanadyl phthalocyanine (VOPc)—on three metallic substrates: gold (Au), silver (Ag), and copper (Cu). This topic can guide new researchers to find out how molecular adsorbance and orientation determine the electronic structure by considering the surface–molecule interactions. Full article
(This article belongs to the Section Applied Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic view of area provided by (<b>a</b>) rough surface, and (<b>b</b>) smooth surface.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic view of the relaxed H<sub>2</sub>Pc molecule on the Au (111) surface: top view and side view. (<b>b</b>) C 1s and (<b>c</b>) N 1s core level spectra for coverage-dependent Au (111)/H<sub>2</sub>Pc hybrid structure. Reprinted (adapted) with permission from [<a href="#B31-molecules-30-00471" class="html-bibr">31</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>Left-hand side</b>) The SPA-LEED pattern of the different phases and (<b>right-hand side</b>) the UPS spectra of 0.7 ML of CuPc and H<sub>2</sub>Pc on Ag(111). Reprinted (abstract/excerpt/figure) with permission from [<a href="#B79-molecules-30-00471" class="html-bibr">79</a>]. Copyright (2024) by the American Physical Society.</p>
Full article ">Figure 4
<p>(<b>Left-hand side</b>) The thickness-dependent evolution of the N 1s core level spectra for Ag (110)/H<sub>2</sub>Pc [A = 7 nm, B–D = 0.4 nm] with the indicated temperature in the panels. (<b>Right-hand side</b>) The STM image for annealed Ag (110)/H<sub>2</sub>Pc: (<b>a</b>) H<sub>2</sub>Pc monolayer (100 °C), and (<b>b</b>) H<sub>2</sub>Pc monolayer (200 °C). Reprinted (adapted) with permission from [<a href="#B84-molecules-30-00471" class="html-bibr">84</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 5
<p>(<b>a</b>) The TPPE spectrum for H<sub>2</sub>Pc monolayer at τ = 0, (<b>b</b>) kinetics trace for the interface state, and (<b>c</b>) the TPPE spectra for the H<sub>2</sub>Pc monolayer at τ = 80 fs with fits, a sloop of one shows a visible. Reprinted (adapted) with permission from [<a href="#B78-molecules-30-00471" class="html-bibr">78</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 6
<p>STM image gathered from the Cu (100)/H<sub>2</sub>Pc structure: (<b>a</b>) typical large area [V<sub>b</sub> = −2.58 V; I<sub>t</sub> = −0.05 nA], (<b>b</b>) two different H<sub>2</sub>Pc molecular orientations [V<sub>b</sub> = −2.58 V; I<sub>t</sub> = −0.04 nA], and (<b>c</b>) modeled top view of the Cu (100)/H<sub>2</sub>Pc structure and the corresponding adsorption configurations for H<sub>2</sub>Pc on the Cu (100) surface. Reprinted (adapted) with permission from [<a href="#B30-molecules-30-00471" class="html-bibr">30</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 7
<p>STM images of the (<b>a</b>) Cu (111)/CuPc hybrid layer [U<sub>bias</sub> = 1.0 V; I<sub>setpoint</sub> = 60 pA] and (<b>b</b>) 0.1 ML of CuPc deposited on the Cu (111) substrate [U<sub>bias</sub> = +0.4 V; I<sub>setpoint</sub> = 20 pA]. The hexagonal lattice of the Cu(111) is shown by yellow lines. Reprinted from [<a href="#B108-molecules-30-00471" class="html-bibr">108</a>], Copyright (2024), with permission from Elsevier. (<b>c</b>) STM images of CuPc molecules deposited on the Cu (100) [Vs = −0.1 V; I = 1 nA; size = 9.0 × 3.9 nm]; (<b>d</b>) CuPc molecules deposited on the Cu (100) utilizing no modified tip [Vs = −0.01 V; I = 5 nA; size = 3.9 × 3.9 nm]; and (<b>e</b>) the modeled view Cu (100)/CuPc hybrid structure. STS graphs for CuPc molecules on (<b>f</b>) Cu (100) and (<b>g</b>) Cu (110); the inset presents the points at which the spectra were collected. Reprinted from [<a href="#B94-molecules-30-00471" class="html-bibr">94</a>], Copyright (2024), with permission from Elsevier.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>b</b>) Two preferred molecular orientations of FePc molecules on the Au (111). Both a and b orientations have been observed experimentally by the high-resolution STM technique. (<b>c</b>) Model of the FePc molecules’ hexamer formed on Au (111) [<a href="#B35-molecules-30-00471" class="html-bibr">35</a>]. STM images of (<b>d</b>) ~0.1 ML (14 nm × 14 nm), (<b>e</b>) ~0.3 ML (14 nm × 14 nm), and (<b>f</b>) ~0.6 ML (14 nm × 14 nm) FePc molecules on the Au(111) surface. Reprinted (adapted) with permission from [<a href="#B35-molecules-30-00471" class="html-bibr">35</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 9
<p>The thickness-dependent evolution of Fe 2p core and shallow HOMO level spectra. Reprinted (adapted) with permission from [<a href="#B120-molecules-30-00471" class="html-bibr">120</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 10
<p>STM images of the Cu (111)/FePc hybrid layer for different overlayers: (<b>a</b>) At thickness much below 1 ML, FePc molecules appear isolated [V = 0.2 V; I = 100.0 pA; size = 23 × 23 nm<sup>2</sup>]. (<b>b</b>) At a larger stage, close but below 1 ML, short-range ordered islands showed up but without any preference (identified by the FFT results presented as inset) [V = 1.6 V; I = 100.0 pA; size = 30 × 30 nm<sup>2</sup>]. (<b>c</b>) [V = 0.29 V; I = 40.0 pA; size = 10 × 10 nm<sup>2</sup>] and (<b>d</b>) [V = 0.6 V; I = 30.0 pA; size = 7.7 × 7.7 nm<sup>2</sup>]) At thickness ~1 ML, the well-ordered domains are obvious. However, based on the FFT results presented as inset, there is no domain preference. Reprinted from [<a href="#B91-molecules-30-00471" class="html-bibr">91</a>]. Copyright (2024), with permission from Elsevier.</p>
Full article ">Figure 11
<p>(<b>a</b>) STM image of CoPc mixed with F<sub>16</sub>CuPc on the Ag(100) [V = −0.78 V and I = 0.25 nA], with the LEED pattern as inset, (<b>b</b>) the HR-STM image of the structure [V = −2.36 V and I = 0.36 nA], and (<b>c</b>) the schematic model of the structure. Reprinted from [<a href="#B139-molecules-30-00471" class="html-bibr">139</a>], with the permission of AIP Publishing.</p>
Full article ">Figure 12
<p>Thickness-dependent XPS and UPS results for Cu (111)/CoPc overlayers: (<b>a</b>) C 1s and (<b>b</b>) N 1s core level spectra. (<b>c</b>) Coverage-dependent valence band obtained for Cu (111)/CoPc. Reprinted (adapted) with permission from [<a href="#B25-molecules-30-00471" class="html-bibr">25</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 13
<p>(<b>a</b>) The schematic of the VOPc adsorbed on the Au (111) substrate. Reprinted from [<a href="#B146-molecules-30-00471" class="html-bibr">146</a>]. Copyright (2024), with permission from Elsevier. (<b>b</b>) High-resolution STM image of the adsorbed VOPc molecule on the Au (111). Reprinted (adapted) with permission from [<a href="#B37-molecules-30-00471" class="html-bibr">37</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">Figure 14
<p>STM images of ~0.2 ML of VOPc molecules on the Cu (111) substrate: (<b>a</b>) Randomly distributed VOPc molecules and some face-down linear chains of VOPcs [V<sub>tip</sub> = 1 V; I = 85 pA; size = 50 × 50 nm<sup>2</sup>]. (<b>b</b>) Face-down linear chains of VOPcs with the same orientation and 1.75 nm of molecular distance [V<sub>tip</sub> = −0.1 V; I = 80 pA; size = 10 × 10 nm<sup>2</sup>]. Enlarged view [V<sub>tip</sub> = 0.1 V; I = 80 pA] of (<b>c</b>) face-up and (<b>d</b>) face-down VOPc molecules. Reprinted (adapted) with permission from [<a href="#B19-molecules-30-00471" class="html-bibr">19</a>]. Copyright {2024} American Chemical Society.</p>
Full article ">
38 pages, 7864 KiB  
Article
An Improved Bio-Inspired Material Generation Algorithm for Engineering Optimization Problems Including PV Source Penetration in Distribution Systems
by Mona Gafar, Shahenda Sarhan, Ahmed R. Ginidi and Abdullah M. Shaheen
Appl. Sci. 2025, 15(2), 603; https://doi.org/10.3390/app15020603 - 9 Jan 2025
Viewed by 559
Abstract
The Material Generation Optimization (MGO) algorithm is an innovative approach inspired by material chemistry which emulates the processes of chemical compound formation and stabilization to thoroughly explore and refine the parameter space. By simulating the bonding processes—such as the formation of ionic and [...] Read more.
The Material Generation Optimization (MGO) algorithm is an innovative approach inspired by material chemistry which emulates the processes of chemical compound formation and stabilization to thoroughly explore and refine the parameter space. By simulating the bonding processes—such as the formation of ionic and covalent bonds—MGO generates new solution candidates and evaluates their stability, guiding the algorithm toward convergence on optimal parameter values. To improve its search efficiency, this paper introduces an Enhanced Material Generation Optimization (IMGO) algorithm, which integrates a Quadratic Interpolated Learner Process (QILP). Unlike conventional random selection, QILP strategically selects three distinct chemical compounds, resulting in increased diversity, a more thorough exploration of the solution space, and improved resistance to local optima. The adaptable and non-linear adjustments of QILP’s quadratic function allow the algorithm to traverse complex landscapes more effectively. This innovative IMGO, along with the original MGO, is developed to support applications across three phases, showcasing its versatility and enhanced optimization capabilities. Initially, both the original and improved MGO algorithms are evaluated using several mathematical benchmarks from the CEC 2017 test suite and benchmarks to measure their optimization capabilities. Following this, both algorithms are applied to the following three well-known engineering optimization problems: the welded beam design, rolling element bearing design, and pressure vessel design. The simulation results are then compared to various established bio-inspired algorithms, including Artificial Ecosystem Optimization (AEO), Fitness–Distance-Balance AEO (FAEO), Chef-Based Optimization Algorithm (CBOA), Beluga Whale Optimization Algorithm (BWOA), Arithmetic-Trigonometric Optimization Algorithm (ATOA), and Atomic Orbital Searching Algorithm (AOSA). Moreover, MGO and IMGO are tested on a real Egyptian power distribution system to optimize the placement of PV and the capacitor units with the aim of minimizing energy losses. Lastly, the PV parameters estimation problem is successfully solved via IMGO, considering the commercial RTC France cell. Comparative studies demonstrate that the IMGO algorithm not only achieves significant energy loss reduction but also contributes to environmental sustainability by reducing emissions, showcasing its overall effectiveness in practical energy optimization applications. The IMGO algorithm improved the optimization outcomes of 23 benchmark models with an average accuracy enhancement of 65.22% and a consistency of 69.57% compared to the MGO method. Also, the application of IMGO in PV parameter estimation achieved a reduction in computational errors of 27.8% while maintaining superior optimization stability compared to alternative methods. Full article
(This article belongs to the Special Issue Heuristic and Evolutionary Algorithms for Engineering Optimization)
Show Figures

Figure 1

Figure 1
<p>Main steps of MGO.</p>
Full article ">Figure 2
<p>Main steps of the proposed IMGO.</p>
Full article ">Figure 3
<p>Visual comparison of key performance indicators between IMGO and MGO for the 23 benchmarks.</p>
Full article ">Figure 4
<p>WBD problem.</p>
Full article ">Figure 5
<p>Best convergence curves of MGO and IMGO for the WBD problem.</p>
Full article ">Figure 6
<p>Heatmap of the obtained cost using MGO and IMGO for the WBD problem.</p>
Full article ">Figure 7
<p>REBD problem.</p>
Full article ">Figure 8
<p>Best convergence curves of MGO and IMGO for the REBD problem.</p>
Full article ">Figure 9
<p>Heatmap of the obtained cost using MGO and IMGO for the REBD problem.</p>
Full article ">Figure 10
<p>PVD problem.</p>
Full article ">Figure 11
<p>Best convergence curves of MGO and IMGO for the PVD problem.</p>
Full article ">Figure 12
<p>Heatmap of the obtained cost using MGO and IMGO for the PVD problem.</p>
Full article ">Figure 12 Cont.
<p>Heatmap of the obtained cost using MGO and IMGO for the PVD problem.</p>
Full article ">Figure 13
<p>Tala distribution feeder.</p>
Full article ">Figure 14
<p>Hourly loading variations.</p>
Full article ">Figure 15
<p>Best convergence curves of MGO, IMGO, DE, PSO, GA, and MVO for energy loss minimization.</p>
Full article ">Figure 16
<p>Heatmap of the obtained cost using MGO and IMGO for energy loss/day (kWday) minimization.</p>
Full article ">Figure 17
<p>Box and whisker plot of MGO, IMGO, DE, PSO, GA, and MVO for energy loss minimization.</p>
Full article ">Figure 18
<p>Schematic diagram of DD-PV model. Where: <span class="html-italic">I<sub>Ph</sub></span> is the irradiation light current. <span class="html-italic">R<sub>sh</sub></span> and <span class="html-italic">R<sub>ss</sub></span> are the parallel and series resistances. <span class="html-italic">η</span><sub>1</sub> and <span class="html-italic">η</span><sub>2</sub> are the ideality factor regarding the two diodes. Id1 and Id2 are the diodes reverse saturation currents.</p>
Full article ">Figure 19
<p>IMGO extracted versus experimental (<b>a</b>) I–V and (<b>b</b>) P-V characteristics for the R.T.C. France cell.</p>
Full article ">Figure 19 Cont.
<p>IMGO extracted versus experimental (<b>a</b>) I–V and (<b>b</b>) P-V characteristics for the R.T.C. France cell.</p>
Full article ">Figure 20
<p>Sensitivity analysis by varying the solutions and iterations for MGO and IMGO for R.T.C. cell parameter estimation.</p>
Full article ">
52 pages, 6163 KiB  
Review
Secondary Metabolites from the Mangrove Ecosystem-Derived Fungi Penicillium spp.: Chemical Diversity and Biological Activity
by Guojun Zhou, Jin Cai, Bin Wang, Wenjiao Diao, Yu Zhong, Shaodan Pan, Weijia Xiong, Guolei Huang and Caijuan Zheng
Mar. Drugs 2025, 23(1), 7; https://doi.org/10.3390/md23010007 - 26 Dec 2024
Viewed by 639
Abstract
Mangrove ecosystems have attracted widespread attention because of their high salinity, muddy or sandy soil, and low pH, as well as being partly anoxic and periodically soaked by tides. Mangrove plants, soil, or sediment-derived fungi, especially the Penicillium species, possess unique metabolic pathways [...] Read more.
Mangrove ecosystems have attracted widespread attention because of their high salinity, muddy or sandy soil, and low pH, as well as being partly anoxic and periodically soaked by tides. Mangrove plants, soil, or sediment-derived fungi, especially the Penicillium species, possess unique metabolic pathways to produce secondary metabolites with novel structures and potent biological activities. This paper reviews the structural diversity and biological activity of secondary metabolites isolated from mangrove ecosystem-derived Penicillium species over the past 5 years (January 2020–October 2024), and 417 natural products (including 170 new compounds, among which 32 new compounds were separated under the guidance of molecular networking and the OSMAC approach) are described. The structures were divided into six major categories, including alkaloids, polyketides, terpenoids, benzene derivatives, steroids, and other classes. Among these natural products, the plausible biosynthetic pathways of 37 compounds were also proposed; 11 compounds have novel skeleton structures, and 26 compounds contain halogen atoms. A total of 126 compounds showed biological activities, such as cytotoxic, antifungal, antibacterial, anti-inflammatory, and α-glucosidase-inhibitory activities, and 11 compounds exhibited diverse biological activities. These new secondary metabolites with novel structures and potent bioactivities will continue to guide the separation or synthesis of structurally novel and biologically active compounds and will offer leading compounds for the development and innovation of pharmaceuticals and pesticides. Full article
(This article belongs to the Special Issue Bioactive Secondary Metabolites of Marine Fungi, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of diketopiperazines from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 2
<p>Chemical structures of indole-diterpenoid alkaloids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 3
<p>Chemical structures of quinolinone, isoquinoline and quinolone alkaloids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 4
<p>Chemical structures of sclerotioramines from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 5
<p>Chemical structures of pyridine derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 6
<p>Chemical structures of benzodiazepines from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 7
<p>Chemical structures of other alkaloids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 8
<p>Chemical structures of lactone derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 8 Cont.
<p>Chemical structures of lactone derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 9
<p>Chemical structures of pyrone derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 10
<p>Chemical structures of azaphilone derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 11
<p>Chemical structures of other polyketides from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 12
<p>Chemical structures of sesquiterpenes from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 13
<p>Chemical structures of meroterpenoids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 13 Cont.
<p>Chemical structures of meroterpenoids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 13 Cont.
<p>Chemical structures of meroterpenoids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 14
<p>Chemical structures of benzene derivatives from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 15
<p>Chemical structures of steroids from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 16
<p>Chemical structures of other classes from mangrove ecosystem-derived <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 17
<p>Structural diversity of the secondary metabolites from the mangrove ecosystem-derived fungus <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 18
<p>Bioactivities of the secondary metabolites from the mangrove ecosystem-derived fungus <span class="html-italic">Penicillium</span> spp.</p>
Full article ">Figure 19
<p>The proportion of <span class="html-italic">Penicillium</span> species from different sources of mangrove ecosystems.</p>
Full article ">Scheme 1
<p>Proposed biosynthetic pathway for compounds <b>39</b>–<b>52</b>.</p>
Full article ">Scheme 2
<p>Proposed biosynthetic pathway for compound <b>135</b>.</p>
Full article ">Scheme 3
<p>Proposed biosynthetic pathway for compounds <b>206</b>–<b>211</b>.</p>
Full article ">Scheme 4
<p>Proposed biosynthetic pathway for compounds <b>278</b>–<b>280</b>.</p>
Full article ">Scheme 5
<p>Proposed biosynthetic pathway for compounds <b>289</b>–<b>291</b>.</p>
Full article ">Scheme 6
<p>Proposed biosynthetic pathway for compounds <b>317</b>–<b>325</b>.</p>
Full article ">Scheme 7
<p>Proposed biosynthetic pathway for compound <b>336</b>.</p>
Full article ">
17 pages, 6429 KiB  
Article
Discovery and Characterization of Two Selective Inhibitors for a Mu-Class Glutathione S-Transferase of 25 kDa from Taenia solium Using Computational and Bioinformatics Tools
by César Sánchez-Juárez, Roberto Flores-López, Lluvia de Carolina Sánchez-Pérez, Ponciano García-Gutiérrez, Lucía Jiménez, Abraham Landa and Rafael A. Zubillaga
Biomolecules 2025, 15(1), 7; https://doi.org/10.3390/biom15010007 - 25 Dec 2024
Viewed by 605
Abstract
Glutathione S-transferases (GSTs) are promising pharmacological targets for developing antiparasitic agents against helminths, as they play a key role in detoxifying cytotoxic xenobiotics and managing oxidative stress. Inhibiting GST activity can compromise parasite viability. This study reports the successful identification of two selective [...] Read more.
Glutathione S-transferases (GSTs) are promising pharmacological targets for developing antiparasitic agents against helminths, as they play a key role in detoxifying cytotoxic xenobiotics and managing oxidative stress. Inhibiting GST activity can compromise parasite viability. This study reports the successful identification of two selective inhibitors for the mu-class glutathione S-transferase of 25 kDa (Ts25GST) from Taenia solium, named i11 and i15, using a computationally guided approach. The workflow involved modeling and refining the 3D structure from the sequence using the AlphaFold algorithm and all-atom molecular dynamics simulations with an explicit solvent. Representative structures from these simulations and a putative binding site with low conservation relative to human GSTs, identified via the SILCS methodology, were employed for virtual screening through ensemble docking against a commercial compound library. The two compounds were found to reduce the enzyme’s activity by 50–70% under assay conditions, while showing a reduction of only 30–35% for human mu-class GSTM1, demonstrating selectivity for Ts25GST. Notable, i11 displayed competitive inhibition with CDNB, while i15 exhibited a non-competitive inhibition type. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Homology model of Ts25GST built using AlphaFold2 and template PDB-ID 1GSU. (<b>A</b>) Front view of the Ts25GST model, highlighting the mu-loops in brown. Chain A is shown in light gray, and Chain B in gray. (<b>B</b>) Molecular surface representation of Ts25GST, illustrating the GSH molecules bound at the G sites. (<b>C</b>) Comparative analysis of the mu-loop from Ts25GST chain A (green) and chain B (black) with three human class Mu M1 GST structures as follows: without GSH (yellow, PDB-ID 1GTU); with GSH bound (brown, PDB-ID 1XW6); and with a GSH-CDNB conjugate (light gray, PDB-ID 1XWK).</p>
Full article ">Figure 2
<p>Molecular dynamics results of Ts25GST-GSH. (<b>A</b>) Normalized PCA; (<b>B</b>) 1D projection of the first five components; (<b>C</b>) free energy landscape projection on the first two components; (<b>D</b>) clustering on energy minima; (<b>E</b>) superposition of the five representative cluster structures; (<b>F</b>) close-up of the G site in the superposition of the 5 conformers; and (<b>G</b>) distance variation between the oxygen of the OH group in Tyr7 and the sulfur of the thiol in GSH.</p>
Full article ">Figure 3
<p>SILCS results from the three replicas (<b>A</b>–<b>C</b>) performed with the solvated model in a mixed solvent of 0.2 M isopropanol/H<sub>2</sub>O. The regions with the highest occupation density of the probe and the estimation of the free energy grid of the points classified as critical according to our methodology are shown.</p>
Full article ">Figure 4
<p>Virtual screening process. (<b>A</b>) Relaxed screening of the complete compound library; (<b>B</b>) exhaustive screening of the top results; and (<b>C</b>) screening on human mu-class GST structures (HGSTM).</p>
Full article ">Figure 5
<p>Results of the purification and enzymatic activity of recombinant Ts25GST. (<b>A</b>) PAGE-SDS of the recombinant protein. (<b>B</b>) Enzyme kinetics with variable CDNB. (<b>C</b>) Enzyme kinetics with variable GSH.</p>
Full article ">Figure 6
<p>Results of the in vitro inhibitory activity assays (<b>A</b>); significance values adjusted by FDR-corrected <span class="html-italic">T</span>-tests [<a href="#B47-biomolecules-15-00007" class="html-bibr">47</a>], selectivity (<b>B</b>); and comparison of the 2D structure and docking results of the two identified inhibitors (<b>C</b>).</p>
Full article ">Figure 7
<p>Experimental results of inhibitor characterization. (<b>A</b>) Dose–response curve for IC<sub>50</sub> calculation. (<b>B</b>,<b>C</b>) Two-dimensional diagrams of the interaction between <span class="html-italic">i11</span> and <span class="html-italic">i15</span> with chain A of the Ts25GST model, respectively. Effect of the concentration of inhibitors <span class="html-italic">i11</span> (<b>D</b>) and <span class="html-italic">i15</span> (<b>E</b>) on the kinetic parameters of recombinant Ts25GST for CDNB variation. Double reciprocal plots for the reaction rates with different concentrations of CDNB and in the presence of different concentrations of inhibitors <span class="html-italic">i11</span> (<b>F</b>) and <span class="html-italic">i15</span> (<b>G</b>).</p>
Full article ">Figure 8
<p>Superposition of conformations of the energy minima from the Ts25GST-GSH simulations in complex with 2 molecules of <span class="html-italic">i11</span> (<b>A</b>) and <span class="html-italic">i15</span> (<b>B</b>). Results of RMSF calculations (<b>C</b>). Analysis of the effect of the inhibitors on the mobility of the mu loop (<b>D</b>).</p>
Full article ">
10 pages, 2377 KiB  
Article
Roughing Nitrogen-Doped Carbon Nanosheets for Loading of Monatomic Fe and Electroreduction of CO2 to CO
by Yuxuan Liu, Yufan Tan, Keyi Zhang, Tianqi Guo, Yao Zhu, Ting Cao, Haiyang Lv, Junpeng Zhu, Ze Gao, Su Zhang, Zheng Liu and Juzhe Liu
Molecules 2024, 29(23), 5561; https://doi.org/10.3390/molecules29235561 - 25 Nov 2024
Viewed by 606
Abstract
The catalyst is the pivotal component in CO2 electroreduction systems for converting CO2 into valuable products. Carbon-based single-atom materials (CSAMs) have emerged as promising catalyst candidates due to their low cost and high atomic utilization efficiency. The rational design of the [...] Read more.
The catalyst is the pivotal component in CO2 electroreduction systems for converting CO2 into valuable products. Carbon-based single-atom materials (CSAMs) have emerged as promising catalyst candidates due to their low cost and high atomic utilization efficiency. The rational design of the morphology and microstructure of such materials is desirable but poses a challenge. Here, we employed different Mg(OH)2 templates to guide the fabrication of two kinds of amorphous nitrogen-doped carbon nanosheet-supported Fe single atoms (FeSNC) with rough and flat surface structures. In comparison to flat FeSNC with saturated FeN4 sites, the rough FeSNC (R-FeSNC) exhibited unsaturated FeN4−x sites and contracted Fe-N bond length. The featured structure endowed R-FeSNC with superior capacity of catalyzing CO2 reduction reaction, achieving an exceptional CO selectivity with Faradaic efficiency of 93% at a potential of −0.66 V vs. RHE. This study offers valuable insights into the design of CSAMs and provides a perspective for gaining a deeper understanding of their activity origins. Full article
(This article belongs to the Special Issue Carbon-Based Electrochemical Materials for Energy Storage)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesis and morphology characterization of R-FeSNC and F-FeSNC. (<b>A</b>) Schematic illustration of the synthetic route for R-FeSNC and F-FeSNC. (<b>B</b>) TEM and (<b>C</b>) HAADF-STEM images of R-FeSNC; the yellow circles mark single metal atoms. (<b>D</b>) Relevant EDS mapping images. (<b>E</b>) TEM and (<b>F</b>) HAADF-STEM images of F-FeSNC; the yellow circles mark single metal atoms. (<b>G</b>) Corresponding EDS mapping images.</p>
Full article ">Figure 2
<p>XAS characterization of R-FeSNC and F-FeSNC. (<b>A</b>) The normalized Fe K-edge XANES and (<b>B</b>) FT-EXAFS spectra of R-FeSNC and F-FeSNC, and the notation "*0.3" appended to "Fe foil" denotes that the peak intensity of the Fe K-edge EXAFS spectrum for the Fe foil has been multiplied by a factor of 0.3. The fitting results of EXAFS spectra for (<b>C</b>) R-FeSNC and (<b>D</b>) F-FeSNC. The wavelet transforms of (<b>E</b>) R-FeSNC and (<b>F</b>) F-FeSNC.</p>
Full article ">Figure 3
<p>XPS studies of R-FeSNC. (<b>A</b>) Fe 2p and (<b>B</b>) N1s XPS spectra of R-FeSNC.</p>
Full article ">Figure 4
<p>Electrocatalytic performance. (<b>A</b>) LSV curves of R-FeSNC and F-FeSNC with CO<sub>2</sub> and Ar feeding. (<b>B</b>) The CO FEs of R-FeSNC and F-FeSNC at different potentials. (<b>C</b>) C<sub>dl</sub> fitting curves. (<b>D</b>) EIS data and equivalent circuit diagram fitting. (<b>E</b>) Stability test of R-FeSNC at a potential of −0.66 V vs. RHE.</p>
Full article ">Figure 5
<p>In situ ATR-FTIR spectra of (<b>A</b>) R-FeSNC and (<b>B</b>) F-FeSNC recorded at different applied potentials for CRR in a CO<sub>2</sub>-saturated 0.5 M KHCO<sub>3</sub> solution. *CO and *COOH are adsorbed intermediates in the carbon dioxide reduction process.</p>
Full article ">
30 pages, 1150 KiB  
Review
Methods for Detection, Extraction, Purification, and Characterization of Exopolysaccharides of Lactic Acid Bacteria—A Systematic Review
by Manoj Kumar Yadav, Ji Hoon Song, Robie Vasquez, Jae Seung Lee, In Ho Kim and Dae-Kyung Kang
Foods 2024, 13(22), 3687; https://doi.org/10.3390/foods13223687 - 19 Nov 2024
Viewed by 2913
Abstract
Exopolysaccharides (EPSs) are large-molecular-weight, complex carbohydrate molecules and extracellularly secreted bio-polymers released by many microorganisms, including lactic acid bacteria (LAB). LAB are well known for their ability to produce a wide range of EPSs, which has received major attention. LAB-EPSs have the potential [...] Read more.
Exopolysaccharides (EPSs) are large-molecular-weight, complex carbohydrate molecules and extracellularly secreted bio-polymers released by many microorganisms, including lactic acid bacteria (LAB). LAB are well known for their ability to produce a wide range of EPSs, which has received major attention. LAB-EPSs have the potential to improve health, and their applications are in the food and pharmaceutical industries. Several methods have been developed and optimized in recent years for producing, extracting, purifying, and characterizing LAB-produced EPSs. The simplest method of evaluating the production of EPSs is to observe morphological features, such as ropy and mucoid appearances of colonies. Ethanol precipitation is widely used to extract the EPSs from the cell-free supernatant and is generally purified using dialysis. The most commonly used method to quantify the carbohydrate content is phenol–sulfuric acid. The structural characteristics of EPSs are identified via Fourier transform infrared, nuclear magnetic resonance, and X-ray diffraction spectroscopy. The molecular weight and composition of monosaccharides are determined through size-exclusion chromatography, thin-layer chromatography, gas chromatography, and high-performance liquid chromatography. The surface morphology of EPSs is observed via scanning electron microscopy and atomic force microscopy, whereas thermal characteristics are determined through thermogravimetry analysis, derivative thermogravimetry, and differential scanning calorimetry. In the present review, we discuss the different existing methods used for the detailed study of LAB-produced EPSs, which provide a comprehensive guide on LAB-EPS preparation, critically evaluating methods, addressing knowledge gaps and key challenges, and offering solutions to enhance reproducibility, scalability, and support for both research and industrial applications. Full article
(This article belongs to the Section Food Microbiology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A schematic diagram of various steps involved in the detection, production, extraction, purification, and characterization of exopolysaccharides (EPSs) of lactic acid bacteria (LAB).</p>
Full article ">Figure 2
<p>A schematic diagram of various methods used for the determination of different structural characteristics of exopolysaccharides (EPSs) of lactic acid bacteria (LAB).</p>
Full article ">
22 pages, 6774 KiB  
Article
Atomistic Modeling of Natural Gas Desulfurization Process Using Task-Specific Deep Eutectic Solvents Supported by Graphene Oxide
by Olzhas Ismagambetov, Nakhypbek Aldiyarov, Nurlan Almas, Irina Irgibaeva, Zhadyra Baitassova, Sergei Piskunov, Anuar Aldongarov and Omirzak Abdirashev
Molecules 2024, 29(22), 5282; https://doi.org/10.3390/molecules29225282 - 8 Nov 2024
Viewed by 637
Abstract
This study employs Density Functional Theory (DFT) calculations and traditional all-atom Molecular Dynamics (MD) simulations to reveal atomistic insights into a task-specific Deep Eutectic Solvent (DES) supported by graphene oxide with the aim of mimicking its application in the natural gas desulfurization process. [...] Read more.
This study employs Density Functional Theory (DFT) calculations and traditional all-atom Molecular Dynamics (MD) simulations to reveal atomistic insights into a task-specific Deep Eutectic Solvent (DES) supported by graphene oxide with the aim of mimicking its application in the natural gas desulfurization process. The DES, composed of N,N,N′,N′-tetramthyl-1,6-hexane diamine acetate (TMHDAAc) and methyldiethanolamine (MDEA) supported by graphene oxide, demonstrates improved efficiency in removing hydrogen sulfide from methane. Optimized structure and HOMO-LUMO orbital analyses reveal the distinct spatial arrangements and interactions between hydrogen sulfide, methane, and DES components, highlighting the efficacy of the DES in facilitating the separation of hydrogen sulfide from methane through DFT calculations. The radial distribution function (RDF) and interaction energies, as determined by traditional all-atom MD simulations, provide insights into the specificity and strength of the interactions between the DES components supported by graphene oxide and hydrogen sulfide. Importantly, the stability of the DES structure supported by graphene oxide is maintained after mixing with the fuel, ensuring its robustness and suitability for prolonged desulfurization processes, as evidenced by traditional all-atom MD simulation results. These findings offer crucial insights into the molecular-level mechanisms underlying the desulfurization of natural gas, guiding the design and optimization of task-specific DESs supported by graphene oxide for sustainable and efficient natural gas purification. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized structures of TMHDAAc, MDEA, and task-specific DES. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; and red signifies oxygen.</p>
Full article ">Figure 2
<p>HOMO-LUMO orbitals of MDEA, TMHDAAc, and task-specific DES. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; and red signifies oxygen. Dark green and dark red indicate the locations of molecular orbitals.</p>
Full article ">Figure 3
<p>Optimized structures of task-specific DES, hydrogen sulfide, methane, and desulfurization process. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; yellow denotes sulfur; and red signifies oxygen.</p>
Full article ">Figure 4
<p>HOMO-LUMO orbitals of desulfurization process in presence of methane, hydrogen sulfide, and task-specific DES. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; yellow denotes sulfur; and red signifies oxygen. Dark green and dark red indicate the locations of molecular orbitals.</p>
Full article ">Figure 5
<p>The RDF between TMHDA and Ac in the pure case and in the presence of MDEA as a DES component.</p>
Full article ">Figure 6
<p>RDF between TMHDA and Ac, MDEA and Ac, and TMHDA and MDEA as DES component.</p>
Full article ">Figure 7
<p>Snapshot of MD simulation of desulfurization process of natural gas using task-specific DES. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; yellow denotes sulfur; and red signifies oxygen.</p>
Full article ">Figure 8
<p>RDFs between hydrogen sulfide and DES components.</p>
Full article ">Figure 9
<p>RDF between DES components during desulfurization of natural gas.</p>
Full article ">Figure 10
<p>Snapshot of MD simulation of desulfurization process of natural gas by task-specific DES supported by graphene oxide. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; yellow denotes sulfur; isosurface with blue color denotes graphene oxide; and red signifies oxygen.</p>
Full article ">Figure 11
<p>RDFs between hydrogen sulfide and DES components supported by graphene oxide.</p>
Full article ">Figure 12
<p>RDFs between hydrogen sulfide and DES components in absence and presence of graphene oxide (GO).</p>
Full article ">Figure 13
<p>RDF between DES components supported by graphene oxide during desulfurization of natural gas.</p>
Full article ">Figure 14
<p>RDF between graphene oxide (GO) and other components during desulfurization of natural gas using DES.</p>
Full article ">Figure 15
<p>(<b>A</b>) [C1-TMHDA], (<b>B</b>) Ac, (<b>C</b>) MDEA, (<b>D</b>) hydrogen sulfide, (<b>E</b>) methane, and (<b>F</b>) graphene oxide structures for our calculations. Color legend: white represents hydrogen; gray indicates carbon; blue denotes nitrogen; yellow denotes sulfur; and red signifies oxygen.</p>
Full article ">
11 pages, 2693 KiB  
Article
Erythrocyte Selenium as a Potential Key Indicator for Selenium Supplementation in Low-Selenium Populations: A Selenium Supplementation Study Based on Wistar Rats
by Cunqi Lv, Ruixiang Wang, Qingyu Zeng, Chen Feng, Guijin Li, Shuxiu Hao, Jiacheng Li, Cheng Wang, Huixin Sun, Linlin Du, Yu Zhang, Xinshu Wang, Tong Wang and Qi Li
Nutrients 2024, 16(22), 3797; https://doi.org/10.3390/nu16223797 - 5 Nov 2024
Viewed by 965
Abstract
Background: Selenium (Se) is an essential trace element for maintaining human health, with significant antioxidant and immunoregulatory functions. Inadequate Se intake may be associated with Keshan disease, Kashin–Beck disease, and hypothyroidism. However, effective indicators for scientifically guiding Se supplementation in Se-deficient populations are [...] Read more.
Background: Selenium (Se) is an essential trace element for maintaining human health, with significant antioxidant and immunoregulatory functions. Inadequate Se intake may be associated with Keshan disease, Kashin–Beck disease, and hypothyroidism. However, effective indicators for scientifically guiding Se supplementation in Se-deficient populations are still lacking. Objectives: This study aims to explore the dynamic distribution of Se across various nutritional biomarkers and major organs in rats through a Se supplementation experiment, as well as the pairwise correlations between them, in order to identify reliable nutritional indicators for evaluating Se levels in the body. Methods: Se levels in hair, blood, and major tissues and organs were determined by atomic fluorescence spectrometry, and glutathione peroxidase (GSH-Px) levels were measured using an ELISA. Results: Se supplementation significantly increased Se levels in rat blood, hair, and major organs, as well as GSH-Px levels in blood. Se primarily accumulated in the liver and kidneys, followed by myocardium, spleen, and muscles. Serum and plasma Se were found to be the best indicators of short-term Se intake, while erythrocyte Se levels showed a stronger correlation with Se levels in tissues and organs, making it a better marker for assessing long-term Se nutritional status compared to hair Se. Conclusions: This study demonstrates the potential of erythrocyte Se levels as an indicator for evaluating long-term Se nutritional status, providing scientific evidence for Se nutritional assessments. Full article
Show Figures

Figure 1

Figure 1
<p>Dietary intake, body weight, and organ coefficients of rats in the control group and the Sodium selenite (SS) group at different time points. Flowchart of the selenium (Se) supplementation experiment in rats (<b>A</b>). Changes in water intake (<b>B</b>), feed intake (<b>C</b>), Se intake from diet (<b>D</b>), body weight (<b>E</b>), liver organ coefficient (<b>F</b>), kidney organ coefficient (<b>G</b>), spleen organ coefficient (<b>H</b>), and heart organ coefficient (<b>I</b>) at different time points. Asterisk indicates statistically significant differences between the SS group and the control group at corresponding time points, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 2
<p>Changes in Se nutritional biomarkers in rats from the control group and SS group at different time points. Changes in dorsal hair Se (<b>A</b>), ventral hair Se (<b>B</b>), whole blood Se (<b>C</b>), serum Se (<b>D</b>), plasma Se (<b>E</b>), erythrocyte Se (<b>F</b>), whole blood GSH-Px (<b>G</b>), serum GSH-Px (<b>H</b>), and plasma GSH-Px (<b>I</b>) levels at different time points in rats. Asterisk indicates statistically significant differences in Se levels and GSH-Px activity between the SS group and the control group at corresponding time points, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Changes in Se levels in various tissues and organs of rats from the control group and SS group at different time points. The Se level changes in liver (<b>A</b>), kidneys (<b>B</b>), spleen (<b>C</b>), myocardium (<b>D</b>), and muscles (<b>E</b>) at different time points in rats. Asterisk indicates statistically significant differences in Se levels between the SS group and the control group at the corresponding time points, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">
17 pages, 6819 KiB  
Review
Theoretical Advances in MBenes for Hydrogen Evolution Electrocatalysis
by Yanwei Wang, Qi Jia, Ge Gao, Ying Zhang, Lei Zhang, Shun Lu and Ling Fang
Energies 2024, 17(21), 5492; https://doi.org/10.3390/en17215492 - 2 Nov 2024
Viewed by 1069
Abstract
Two-dimensional transition metal borides (MBenes) have emerged as promising electrocatalysts for hydrogen evolution reactions (HERs), attracting significant research interest due to theoretical computations that enhance the understanding and optimization of their performance. This review begins with a comprehensive summary of HER mechanisms, followed [...] Read more.
Two-dimensional transition metal borides (MBenes) have emerged as promising electrocatalysts for hydrogen evolution reactions (HERs), attracting significant research interest due to theoretical computations that enhance the understanding and optimization of their performance. This review begins with a comprehensive summary of HER mechanisms, followed by an in-depth examination of the geometric and electronic properties of MBenes. Subsequently, this review explores MBene-based electrocatalysts for HERs, employing free-energy diagrams and an electronic structure analysis to assess both the intrinsic catalytic activity of MBenes and the theoretical performance of single-atom modified MBenes. Finally, the prospects and challenges associated with MBenes are discussed, providing valuable insights to guide future research in this area. Overall, this topic holds significant relevance for researchers in the HER field, and this review aims to deliver theoretical insights for the optimal design of advanced MBene electrocatalysts. Full article
(This article belongs to the Section D: Energy Storage and Application)
Show Figures

Figure 1

Figure 1
<p>Schematic depiction of the hydrogen evolution reaction (HER) mechanism in (<b>a</b>) acidic conditions and (<b>b</b>) alkaline conditions.</p>
Full article ">Figure 2
<p>Volcano curve of HER activity. Reproduced with permission from [<a href="#B29-energies-17-05492" class="html-bibr">29</a>], Springer Nature, 2006.</p>
Full article ">Figure 3
<p>(<b>a</b>) The composition of elements in the MAB phase. Reproduced with permission from [<a href="#B32-energies-17-05492" class="html-bibr">32</a>], Wiley-VCH, 2022. (<b>b</b>) Configurations of MAB phases in orthorhombic and hexagonal forms. Reproduced with permission from [<a href="#B34-energies-17-05492" class="html-bibr">34</a>], Royal Society of Chemistry, 2022. (<b>c</b>) Depictions of elements capable of forming MBenes. Reproduced with permission from [<a href="#B18-energies-17-05492" class="html-bibr">18</a>], American Chemical Society, 2021.</p>
Full article ">Figure 4
<p>(<b>a</b>) Band structure and partial density of states (PDOS) for a TiB single-layer. (<b>b</b>) ELF maps for the TiB single layer and [TiB]<sup>+</sup>. Reproduced with permission from [<a href="#B47-energies-17-05492" class="html-bibr">47</a>], American Chemical Society, 2019. (<b>c</b>) 3D isosurfaces of the electron localization function (ELF) for <span class="html-italic">h</span>-Nb<sub>2</sub>B and <span class="html-italic">h</span>-NbB, both with and without O functionalization, plotted at the isosurface value of 0.75. (<b>d</b>) Band structures of <span class="html-italic">h</span>-Nb<sub>2</sub>B and <span class="html-italic">h</span>-Nb<sub>2</sub>BO<sub>2</sub>. (<b>e</b>) Band structures of <span class="html-italic">h</span>-NbB and <span class="html-italic">h</span>-NbBO. Reproduced with permission from [<a href="#B49-energies-17-05492" class="html-bibr">49</a>], American Chemical Society, 2024.</p>
Full article ">Figure 5
<p>(<b>a</b>) Computed density of states for 2D Mo<sub>2</sub>B<sub>2</sub>. (<b>b</b>–<b>d</b>) Band structures for 2D Mo<sub>2</sub>B<sub>2</sub> and Fe<sub>2</sub>B<sub>2</sub>, where the sizes of the red/blue circles and green squares indicate the projected contributions of Mo/Fe d orbitals and B atom p orbitals, respectively. Reproduced with permission from [<a href="#B53-energies-17-05492" class="html-bibr">53</a>], Royal Society of Chemistry, 2017.</p>
Full article ">Figure 6
<p>Calculated H adsorption Gibbs free energy ΔG<sub>H*</sub> of (<b>a</b>) Fe<sub>2</sub>B<sub>2</sub> and (<b>b</b>) Fe<sub>2</sub>B<sub>2</sub>O<sub>2</sub> under varying hydrogen coverages. Reproduced with permission from [<a href="#B53-energies-17-05492" class="html-bibr">53</a>], Royal Society of Chemisry, 2017. The free-energy diagram for the HER at different hydrogen coverages on the surfaces of (<b>c</b>) Cr<sub>2</sub>B<sub>2</sub>, (<b>d</b>) Cr<sub>3</sub>B<sub>4</sub>, and (<b>e</b>) Cr<sub>4</sub>B<sub>6</sub>. Reproduced with permission from [<a href="#B55-energies-17-05492" class="html-bibr">55</a>], Elsevier, 2020. (<b>f</b>) The ΔG<sub>H*</sub> of Mn<sub>2</sub>B<sub>2</sub> under varying hydrogen coverage. Reproduced with permission from [<a href="#B56-energies-17-05492" class="html-bibr">56</a>], Elsevier, 2020. The ΔG<sub>H*</sub> with different hydrogen coverage on (<b>g</b>) ZrBO and (<b>h</b>) Nb<sub>2</sub>BO<sub>2</sub>. Reproduced with permission from [<a href="#B49-energies-17-05492" class="html-bibr">49</a>], American Chemical Society, 2024.</p>
Full article ">Figure 7
<p>The ΔG<sub>H*</sub> of (<b>a</b>) M<sub>2</sub>B<sub>2</sub>, (<b>b</b>) M<sub>3</sub>B<sub>4</sub>, and (<b>c</b>) M<sub>4</sub>B<sub>6</sub> at varying hydrogen coverages; atomic configurations of (<b>d</b>) Fe<sub>2</sub>B<sub>2</sub>, (<b>e</b>) Fe<sub>3</sub>B<sub>4</sub>, and (<b>f</b>) Fe<sub>4</sub>B<sub>6</sub> at their optimal adsorption sites under different H coverage levels. Reproduced with permission from [<a href="#B54-energies-17-05492" class="html-bibr">54</a>], Elsevier, 2023.</p>
Full article ">Figure 8
<p>The ΔG<sub>H*</sub> values following the introduction of different metal single atoms into (<b>a</b>) Mo<sub>2</sub>B<sub>2</sub>O<sub>2</sub> and (<b>b</b>) W<sub>2</sub>B<sub>2</sub>O<sub>2</sub>. (<b>c</b>) Three-dimensional structure of V-W<sub>2</sub>B<sub>2</sub>O<sub>2</sub>. Reproduced with permission from [<a href="#B64-energies-17-05492" class="html-bibr">64</a>], American Chemical Society, 2020. The ΔG<sub>H*</sub> of (<b>d</b>) pristine Ti<sub>2</sub>B and (<b>e</b>) TM-Ti<sub>2</sub>B. Reproduced with permission from [<a href="#B48-energies-17-05492" class="html-bibr">48</a>], Elsevier, 2024.</p>
Full article ">Figure 9
<p>Free energy of H* intermediates adsorbed on Pt (111), Pt-MoS<sub>2</sub>, Pt-Mo<sub>2</sub>C, and Pt-MoAl<sub>1-x</sub>B in (<b>a</b>) alkaline and (<b>b</b>) acidic environments. Reproduced with permission from [<a href="#B68-energies-17-05492" class="html-bibr">68</a>], Royal Society of Chemistry, 2023. (<b>c</b>) Gibbs free energy profile for water dissociation on the (100) surface of Ni-MoB<sub>2</sub>. (<b>d</b>) The ΔG<sub>H*</sub> of Ni-MoB<sub>2</sub> on various sites. Reproduced with permission from [<a href="#B69-energies-17-05492" class="html-bibr">69</a>], American Chemical Society, 2023.</p>
Full article ">
14 pages, 3865 KiB  
Article
Assessment of Classical Force-Fields for Graphene Mechanics
by Zhiwei Ma, Yongkang Tan, Xintian Cai, Xue Chen, Tan Shi, Jianfeng Jin, Yifang Ouyang and Qing Peng
Crystals 2024, 14(11), 960; https://doi.org/10.3390/cryst14110960 - 2 Nov 2024
Cited by 1 | Viewed by 931
Abstract
The unique properties of graphene have attracted the interest of researchers from various fields, and the discovery of graphene has sparked a revolution in materials science, specifically in the field of two-dimensional materials. However, graphene synthesis’s costly and complex process significantly impairs researchers’ [...] Read more.
The unique properties of graphene have attracted the interest of researchers from various fields, and the discovery of graphene has sparked a revolution in materials science, specifically in the field of two-dimensional materials. However, graphene synthesis’s costly and complex process significantly impairs researchers’ endeavors to explore its properties and structure experimentally. Molecular dynamics simulation is a well-established and useful tool for investigating graphene’s atomic structure and dynamic behavior at the nanoscale without requiring expensive and complex experiments. The accuracy of the molecular dynamics simulation depends on the potential functions. This work assesses the performance of various potential functions available for graphene in mechanical properties prediction. The following two cases are considered: pristine graphene and pre-cracked graphene. The most popular fifteen potentials have been assessed. Our results suggest that diverse potentials are suitable for various applications. REBO and Tersoff potentials are the best for simulating monolayer pristine graphene, and the MEAM and the AIREBO-m potentials are recommended for those with crack defects because of their respective utilization of the electron density and inclusion of the long-range interaction. We recommend the AIREBO-m potential for a general case of classical molecular dynamics study. This work might help to guide the selection of potentials for graphene simulations and the development of further advanced interatomic potentials. Full article
(This article belongs to the Section Inorganic Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>The different unit cells for graphene. (<b>a</b>) The primitive unitcell that contains only two atoms. The bond length of atom 1 and atom 2 is 0.142 nm. The bond angle is 120°; (<b>b</b>) Conventional unitcell that contains 4 atoms numbered 1–4; (<b>c</b>) A supercell of monolayer of pristine graphene.</p>
Full article ">Figure 2
<p>The stress–strain curve for graphene with different potentials (CH.airebo, CH.rebo, C.meam, CCu, C.lcbop, BNC, SiCGe, SiC-a, SiC-b, SiC-c, SiC-d, SiC-e, SiC-gw, FeC, and SiC-f).</p>
Full article ">Figure 3
<p>Comparison between the simulation results for different interatomic potentials in terms of the mechanical properties. <b>Upper panel</b>: the fracture strain and toughness; <b>Bottom panel</b>: Young’s modulus and failure strength. Blue, grey, black and red bar represent fracture strain, toughness, Young’s modulus, and failure strength, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) The pre-cracked graphene sheet of a size of 20.5 nm × 20.2 nm and 15,981 atoms. (<b>b</b>) The material before making a cut or crack (bulk material). (<b>c</b>) The material after cutting the material and creating two surfaces.</p>
Full article ">Figure 5
<p>The predicted stress intensity factor (SIF) as a function of the half initial length of the crack <span class="html-italic">a</span><sub>0</sub> for the five different potentials compared to experiment.</p>
Full article ">Figure 6
<p>The comparison of the final stress intensity factor (SIF) value and surface energy of the five potentials. Black bar and red bar represents surface energy and final SIF, respectively.</p>
Full article ">
29 pages, 9314 KiB  
Review
Bridging Materials and Analytics: A Comprehensive Review of Characterization Approaches in Metal-Based Solid-State Hydrogen Storage
by Yaohui Xu, Yang Zhou, Yuting Li and Yang Zheng
Molecules 2024, 29(21), 5014; https://doi.org/10.3390/molecules29215014 - 23 Oct 2024
Viewed by 1328
Abstract
The advancement of solid-state hydrogen storage materials is critical for the realization of a sustainable hydrogen economy. This comprehensive review elucidates the state-of-the-art characterization techniques employed in solid-state hydrogen storage research, emphasizing their principles, advantages, limitations, and synergistic applications. We critically analyze conventional [...] Read more.
The advancement of solid-state hydrogen storage materials is critical for the realization of a sustainable hydrogen economy. This comprehensive review elucidates the state-of-the-art characterization techniques employed in solid-state hydrogen storage research, emphasizing their principles, advantages, limitations, and synergistic applications. We critically analyze conventional methods such as the Sieverts technique, gravimetric analysis, and secondary ion mass spectrometry (SIMS), alongside composite and structure approaches including Raman spectroscopy, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM). This review highlights the crucial role of in situ and operando characterization in unraveling the complex mechanisms of hydrogen sorption and desorption. We address the challenges associated with characterizing metal-based solid-state hydrogen storage materials discussing innovative strategies to overcome these obstacles. Furthermore, we explore the integration of advanced computational modeling and data-driven approaches with experimental techniques to enhance our understanding of hydrogen–material interactions at the atomic and molecular levels. This paper also provides a critical assessment of the practical considerations in characterization, including equipment accessibility, sample preparation protocols, and cost-effectiveness. By synthesizing recent advancements and identifying key research directions, this review aims to guide future efforts in the development and optimization of high-performance solid-state hydrogen storage materials, ultimately contributing to the broader goal of sustainable energy systems. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of a Sieverts apparatus [<a href="#B34-molecules-29-05014" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>(<b>a</b>) Diagram of temperature gradients in the reactor [<a href="#B38-molecules-29-05014" class="html-bibr">38</a>]. The MgH<sub>2</sub> + 7 wt.% Ni/VN (<b>b</b>) PCT curves at 325 and 275 °C and (<b>c</b>) Van’t Hoff plots [<a href="#B72-molecules-29-05014" class="html-bibr">72</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) The DSC profiles at 2, 5, 8, and 10 °C·min<sup>−1</sup> of MgH<sub>2</sub>-7CeZrO and ball-milled MgH<sub>2</sub>; (<b>b</b>) the apparent dehydrogenation activation energy of MgH<sub>2</sub>-7CeZrO and undoped MgH<sub>2</sub> by Kissinger equation.</p>
Full article ">Figure 4
<p>A schematic overview of the secondary ion mass spectrometry experiment [<a href="#B51-molecules-29-05014" class="html-bibr">51</a>]. (<b>a</b>) A surface is bombarded with a primary ion resulting in the sputtering of secondary ions characteristic of surface chemistry. Secondary ions are detected and measured by mass spectrometry. The bombardment is by primary ions, ranging from atomic ions offering the highest lateral resolution to massive gas cluster ion beams that liberate surface species up to several thousand mass units. (<b>b</b>) Mass analysis of secondary ions is generally by quadrupole magnetic sector, time-of-flight, or Orbitrap instruments. (<b>c</b>) Outputs from the analysis include mass spectra, 2D or 3D images, and depth profiles, which can be further processed using machine learning. EM, electromagnetic; MCP, microchannel plate [<a href="#B51-molecules-29-05014" class="html-bibr">51</a>].</p>
Full article ">Figure 5
<p>SIMS images from a hydrogenated sample with (<b>a</b>) equiaxed microstructure (<span class="html-italic">Batch A—Sample 1</span>) showing the summed signal over 3 slices and (<b>b</b>) columnar microstructure (<span class="html-italic">Batch B—Sample 2</span>) showing the summed signal over 14 slices. SIMS localized depth profiles from regions inside and outside of the surface-visible hydride areas for (<b>c</b>) <span class="html-italic">Sample 1</span> and (<b>d</b>) <span class="html-italic">Sample 2</span>. Note that for (<b>d</b>), the “hydride” data have been shifted to correct its depth so that both curves agree on the depth of the substrate (zero on the <span class="html-italic">x</span>-axis), since the hydrides stick above the surface (discussed below). For (<b>a</b>,<b>c</b>), 15 total slices were captured with an average slice thickness of 33 nm. Beam current of 50 pA and a dwell time of 4 ms per pixel. Image resolution of 256 × 256 pixels for FOV of 17 × 17 μm. For (<b>b</b>,<b>d</b>), 56 slices total were collected with an average slice thickness of 14.3 nm. Beam current of 100 pA and a dwell time of 1 ms per pixel. Image resolution of 256 × 256 pixels for FOV of 17 × 17 μm [<a href="#B53-molecules-29-05014" class="html-bibr">53</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic of different configurations of laser excitation and Raman scattered light collection. (<b>b</b>) Schematic of laser excitation and Raman scattered light collection in free space [<a href="#B76-molecules-29-05014" class="html-bibr">76</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) Representative Raman spectra. (<b>b</b>) In situ Raman profile of materials under heating conditions (from room temperature to 200 °C) and (<b>c</b>) simultaneous mass spectrometry profiles for H<sub>2</sub> and other volatile components evolved during thermal decomposition of neat AB under a ramp of 1 °C·min<sup>−1</sup> [<a href="#B82-molecules-29-05014" class="html-bibr">82</a>].</p>
Full article ">Figure 8
<p>FTIR spectra of (a) the commercially purchased bulk KBr powder, (b) hand-mixed MgH<sub>2</sub> + 5 vol% C, (c) BMAS powder, (d) BMAS powder after one dehydrogenation (1R) powder, (e) BMAS powder after one dehydrogenation and then re-hydrogenation (1S) powder, and (f) BMAS powder after 7 cycles of dehydrogenation and re-hydrogenation and then dehydrogenation again (8R) powder [<a href="#B37-molecules-29-05014" class="html-bibr">37</a>].</p>
Full article ">Figure 9
<p>Schematic diagram of the XRD principle.</p>
Full article ">Figure 10
<p>In situ SR-XRD and phase content at 400 °C during (<b>a</b>) hydrogen desorption in dynamic vacuum and (<b>b</b>) cycling between 50 bar H<sub>2</sub> and dynamic vacuum [<a href="#B98-molecules-29-05014" class="html-bibr">98</a>].</p>
Full article ">Figure 11
<p>Schematic diagram of a neutron scattering device [<a href="#B102-molecules-29-05014" class="html-bibr">102</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) X-ray and (<b>b</b>) neutron diffraction patterns of deuterated (1−x)MgD<sub>2</sub>−xTiD<sub>2</sub> nanocomposites for x = 0, 0.1, 0.3, and 0.5 [<a href="#B107-molecules-29-05014" class="html-bibr">107</a>].</p>
Full article ">Figure 13
<p>Schematic diagram of the XPS principle. <span class="html-italic">E</span> is the binding energy [<a href="#B113-molecules-29-05014" class="html-bibr">113</a>].</p>
Full article ">Figure 14
<p>High-resolution XPS spectra of (<b>a</b>) Ti 2p and (<b>b</b>) O 1s, as well as (<b>c</b>) valence changes during the hydrogenation and dehydrogenation processes.</p>
Full article ">Figure 15
<p>SEM analyses of the (<b>a</b>–<b>c</b>) pristine sample, (<b>d</b>–<b>f</b>) aged + PLA sample, (<b>g</b>–<b>i</b>) pristine sample after hydrogenation, and (<b>j</b>–<b>l</b>) aged + PLA sample after hydrogenation [<a href="#B124-molecules-29-05014" class="html-bibr">124</a>].</p>
Full article ">Figure 16
<p>In situ TEM analysis of the hydrogenated MgH<sub>2</sub>/Ni@pCNF composites: (<b>a</b>) HAADF image (the square marked by red dotted line indicates the irradiated area). (<b>b</b>) BF image. (<b>c</b>) The corresponding elemental mapping of C, N, Mg, and Ni. (<b>d</b>–<b>g</b>) HRTEM images and selective electron diffraction at random points showing the evolution of microstructure upon hydrogen desorption induced by the electron-beam irradiation. (<b>d1</b>–<b>d3</b>) Initial microstructure showing lattice fringes of MgH<sub>2</sub> (101), Mg<sub>2</sub>NiH<sub>4</sub> (311), and MgO (200), respectively, before irradiation. (<b>e1</b>–<b>e4</b>) After 3 min, partial decomposition of Mg<sub>2</sub>NiH<sub>4</sub> into Mg<sub>2</sub>Ni begins, with defects forming at the Mg<sub>2</sub>NiH<sub>4</sub>/MgH<sub>2</sub> interface, promoting hydrogen desorption, while some MgH<sub>2</sub> remains stable. (<b>f1</b>–<b>f3</b>) At 6 min, complete decomposition of Mg<sub>2</sub>NiH<sub>4</sub> is observed, while MgH<sub>2</sub> remains partially stable, and Mg nanoparticles become visible. (<b>g1</b>,<b>g2</b>) After 10 min, hydrogen is fully released and transferred to Mg and Mg<sub>2</sub>Ni [<a href="#B133-molecules-29-05014" class="html-bibr">133</a>].</p>
Full article ">Figure 17
<p>AFM micrographs of Ta/Mg/CrV/Pd and Ta/Mg-10%Cr-10%V/CrV/Pd in the as-deposited and hydrogenated state. The Ta/Mg/CrV/Pd was hydrogenated at 50 mbar for 14 h and Ta/Mg-10%Cr-10%V/CrV/Pd at 10 mbar for 20 h. The inset shows the micrograph of the hydrided film on the same brightness scale as the as-deposited state for Ta/Mg/CrV/Pd [<a href="#B139-molecules-29-05014" class="html-bibr">139</a>].</p>
Full article ">
18 pages, 16776 KiB  
Article
Molecular Dynamics Analysis of Multi-Factor Influences on Structural Defects in Deposited Mg-Matrix Zn Atom Films
by Zhen Zhou, Chaoyue Ji, Dongyang Hou, Shunyong Jiang, Yuhang Ouyang, Fang Dong and Sheng Liu
Materials 2024, 17(19), 4700; https://doi.org/10.3390/ma17194700 - 25 Sep 2024
Viewed by 1029
Abstract
Mg metal vascular stents not only have good mechanical support properties but also can be entirely absorbed by the human body as a trace element beneficial to the human body, but because Mg metal is quickly dissolved and absorbed in the human body, [...] Read more.
Mg metal vascular stents not only have good mechanical support properties but also can be entirely absorbed by the human body as a trace element beneficial to the human body, but because Mg metal is quickly dissolved and absorbed in the human body, magnesium metal alone cannot be ideally used as a vascular stent. Since the dense oxide Zn film formed by Zn contact with oxygen in the air has good anti-corrosion performance, the Zn nanolayer film deposited on the Mg matrix vascular scaffold by magnetron sputtering can effectively inhibit the rapid dissolution of Mg metal. However, there are few studies on the molecular dynamic structural defects of Mg-matrix Zn atomic magnetron sputtering, and the atomic level simulation of Mg-matrix Zn thin-film depositions can comprehensively understand the atomic level structural defects in the deposition process of Zn thin films from a theoretical perspective to further guide experimental research. Based on this, this research first studied and analyzed the atomic layer structure defects, surface morphology, surface roughness, atomic density of different deposited layers, radial distribution function, and residual stress of the thin-film deposition layer of 1500 deposited Zn atoms at the initial deposition stage, during and after deposition under Mg-matrix thermal layer 500K and a deposited velocity 5 Å/ps by molecular dynamics. At the same time, the effects of temperature and deposited velocity of the Mg-matrix thermal layer on the surface morphology, roughness, and biaxial stress of the deposited films were studied. Full article
(This article belongs to the Special Issue Advancements in Thin Film Deposition Technologies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mg-matrix Zn deposition model.</p>
Full article ">Figure 2
<p>The curve of potential energy and heating temperature.</p>
Full article ">Figure 3
<p>The molecule dynamic for melting results in Mg and Zn.</p>
Full article ">Figure 4
<p>Different atom coverage for different layers at a thermal part temperature of 500 K and the deposited velocity of 5 Å/ps. (<b>a</b>) The content for Mg and Zn coverage for various layers. (<b>b</b>) Atoms of different functional layers covered for other layers.</p>
Full article ">Figure 5
<p>Time-dependent snapshots of Zn deposited on Mg substrate at a thermal part temperature of 500 K and a deposited velocity of 5 Å/ps.</p>
Full article ">Figure 6
<p>Atomic layer structure defects for different layers of the Mg matrix.</p>
Full article ">Figure 7
<p>Radial distribution function (RDF) of deposited layers with Zn under 500 K.</p>
Full article ">Figure 8
<p>Various structure fractions at a thermal part temperature of 500 K and a deposited velocity of 5 Å/ps. (<b>a</b>) The beginning of the deposition. (<b>b</b>) The end of the deposition.</p>
Full article ">Figure 9
<p>Surface morphology at a thermal part temperature of 500 K and a deposited velocity of 5 Å/ps. (<b>a</b>) Atomic morphology. (<b>b</b>) Surface morphology of the deposited film.</p>
Full article ">Figure 10
<p>The average biaxial stress with the increase in step time at a thermal part temperature of 500 K and a deposited velocity of 5 Å/ps.</p>
Full article ">Figure 11
<p>The atomic layer density of the deposited film with different layer numbers at a thermal part temperature of 500 K and a deposited velocity of 5 Å/ps.</p>
Full article ">Figure 12
<p>Atomic morphology (<b>a</b>–<b>d</b>) and surface morphology (<b>e</b>–<b>h</b>) of the deposited film at different thermal temperatures and a deposited velocity of 5 Å/ps: (<b>a</b>) and (<b>e</b>) at 300 K; (<b>b</b>) and (<b>f</b>) at 500 K; (<b>c</b>) and (<b>g</b>) at 700 K; (<b>d</b>) and (<b>h</b>) 900 K.</p>
Full article ">Figure 12 Cont.
<p>Atomic morphology (<b>a</b>–<b>d</b>) and surface morphology (<b>e</b>–<b>h</b>) of the deposited film at different thermal temperatures and a deposited velocity of 5 Å/ps: (<b>a</b>) and (<b>e</b>) at 300 K; (<b>b</b>) and (<b>f</b>) at 500 K; (<b>c</b>) and (<b>g</b>) at 700 K; (<b>d</b>) and (<b>h</b>) 900 K.</p>
Full article ">Figure 13
<p>The surface roughness of deposited film at different temperatures.</p>
Full article ">Figure 14
<p>The atomic layer density of the deposited film at different thermal temperatures and the deposited velocity of 5 Å/ps.</p>
Full article ">Figure 15
<p>The average biaxial stress with the different step time.</p>
Full article ">Figure 16
<p>Atomic morphology (<b>a</b>–<b>d</b>) and surface morphology (<b>e</b>–<b>h</b>) of the deposited film at different deposited velocities: (<b>a</b>) and (<b>e</b>) at 5 Å/ps; (<b>b</b>) and (<b>f</b>) at 10 Å/ps; (<b>c</b>) and (<b>g</b>) at 15 Å/ps; (<b>d</b>) and (<b>h</b>) 20 Å/ps.</p>
Full article ">Figure 17
<p>The surface roughness of deposited film at different velocities.</p>
Full article ">Figure 18
<p>The atomic layer density of the deposited film with different layer numbers under different velocities under a temperature of 500 K.</p>
Full article ">Figure 19
<p>Relationship curves of biaxial stress and step times at different deposited velocities under a temperature of 500K.</p>
Full article ">
10 pages, 1994 KiB  
Article
Enhanced Thermal Stability of Conductive Mercury Telluride Colloidal Quantum Dot Thin Films Using Atomic Layer Deposition
by Edward W. Malachosky, Matthew M. Ackerman and Liliana Stan
Nanomaterials 2024, 14(16), 1354; https://doi.org/10.3390/nano14161354 - 16 Aug 2024
Cited by 1 | Viewed by 1130
Abstract
Colloidal quantum dots (CQDs) are valuable for their potential applications in optoelectronic devices. However, they are susceptible to thermal degradation during processing and while in use. Mitigating thermally induced sintering, which leads to absorption spectrum broadening and undesirable changes to thin film electrical [...] Read more.
Colloidal quantum dots (CQDs) are valuable for their potential applications in optoelectronic devices. However, they are susceptible to thermal degradation during processing and while in use. Mitigating thermally induced sintering, which leads to absorption spectrum broadening and undesirable changes to thin film electrical properties, is necessary for the reliable design and manufacture of CQD-based optoelectronics. Here, low-temperature metal–oxide atomic layer deposition (ALD) was investigated as a method for mitigating sintering while preserving the optoelectronic properties of mercury telluride (HgTe) CQD films. ALD-coated films are subjected to temperatures up to 160 °C for up to 5 h and alumina (Al2O3) is found to be most effective at preserving the optical properties, demonstrating the feasibility of metal–oxide in-filling to protect against sintering. HgTe CQD film electrical properties were investigated before and after alumina ALD in-filling, which was found to increase the p-type doping and hole mobility of the films. The magnitude of these effects depended on the conditions used to prepare the HgTe CQDs. With further investigation into the interaction effects of CQD and ALD process factors, these results may be used to guide the design of CQD–ALD materials for their practical integration into useful optoelectronic devices. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Infrared absorption spectra plotted against the wavelength (nm) of HgTe CQD thin films on sapphire substrates measured as a function of atomic layer deposition cycles and bake conditions. Evolution of the infrared absorption spectrum for HgTe CQD after (<b>a</b>) ZnO, (<b>b</b>) TiO<sub>2</sub>, (<b>c</b>) 4 cycles of alumina, (<b>d</b>) 8 cycles of alumina, and (<b>e</b>) 20 cycles of alumina and being subjected to baking up to 165 °C for up to 5 h.</p>
Full article ">Figure 2
<p>Log plots of the current versus voltage for HgTe CQD thin film Au-Au photoconductor devices with a 5-microns electrode gap. The conductance for a control device (black) that was not subjected to atomic layer deposition, a film subjected to 8 cycles of alumina ALD (red), and a film subjected to 20 cycles alumina ALD (green) are plotted for comparison. Conductance of HgTe CQD films measured (<b>a</b>) before and (<b>b</b>) after baking at 130 °C for 2 h under a nitrogen environment are shown.</p>
Full article ">Figure 3
<p>Transfer curves for HgTe colloidal quantum dot thin films measured at a 1 V source–drain bias (<b>a</b>) before alumina ALD and (<b>b</b>) after alumina ALD. Black dashed lines indicate the maximum and minimum slopes taken to calculate the carrier mobilities.</p>
Full article ">Figure 4
<p>Transfer curves for n-type doped HgTe colloidal quantum dot thin films measured at a 1 V source–drain bias (<b>a</b>) before alumina ALD and (<b>b</b>) after alumina ALD. Black dashed lines indicate the maximum and minimum slopes taken to calculate the carrier mobilities.</p>
Full article ">
20 pages, 1948 KiB  
Review
Crystal Structure Prediction and Performance Assessment of Hydrogen Storage Materials: Insights from Computational Materials Science
by Xi Yang, Yuting Li, Yitao Liu, Qian Li, Tingna Yang and Hongxing Jia
Energies 2024, 17(14), 3591; https://doi.org/10.3390/en17143591 - 22 Jul 2024
Cited by 2 | Viewed by 1570
Abstract
Hydrogen storage materials play a pivotal role in the development of a sustainable hydrogen economy. However, the discovery and optimization of high-performance storage materials remain a significant challenge due to the complex interplay of structural, thermodynamic and kinetic factors. Computational materials science has [...] Read more.
Hydrogen storage materials play a pivotal role in the development of a sustainable hydrogen economy. However, the discovery and optimization of high-performance storage materials remain a significant challenge due to the complex interplay of structural, thermodynamic and kinetic factors. Computational materials science has emerged as a powerful tool to accelerate the design and development of novel hydrogen storage materials by providing atomic-level insights into the storage mechanisms and guiding experimental efforts. In this comprehensive review, we discuss the recent advances in crystal structure prediction and performance assessment of hydrogen storage materials from a computational perspective. We highlight the applications of state-of-the-art computational methods, including density functional theory (DFT), molecular dynamics (MD) simulations, and machine learning (ML) techniques, in screening, evaluating, and optimizing storage materials. Special emphasis is placed on the prediction of stable crystal structures, assessment of thermodynamic and kinetic properties, and high-throughput screening of material space. Furthermore, we discuss the importance of multiscale modeling approaches that bridge different length and time scales, providing a holistic understanding of the storage processes. The synergistic integration of computational and experimental studies is also highlighted, with a focus on experimental validation and collaborative material discovery. Finally, we present an outlook on the future directions of computationally driven materials design for hydrogen storage applications, discussing the challenges, opportunities, and strategies for accelerating the development of high-performance storage materials. This review aims to provide a comprehensive and up-to-date account of the field, stimulating further research efforts to leverage computational methods to unlock the full potential of hydrogen storage materials. Full article
Show Figures

Figure 1

Figure 1
<p>Statistics on the number of research papers published in the field of solid-state hydrogen storage (as of the end of 2023).</p>
Full article ">Figure 2
<p>(<b>a</b>) DFT-predicted crystal structure and hydrogen adsorption sites in a MOF material. Reproduced with permission from [<a href="#B46-energies-17-03591" class="html-bibr">46</a>], Elsevier, 2022. (<b>b</b>) The model of MOF-519-X and structures of organic linker. Atom color scheme: C, gray; H, white; O, red; X, yellow (X = single bond OH, single bond NO<sub>2</sub>, single bond Cl, single bond NH<sub>2</sub>, single bond CH<sub>3</sub>, single bond F). Reproduced with permission from [<a href="#B47-energies-17-03591" class="html-bibr">47</a>], Elsevier, 2019.</p>
Full article ">Figure 3
<p>MD-simulated diffusion pathways of hydrogen atoms (yellow spheres) in a MgH<sub>2</sub> crystal. The magnesium atoms are shown as green spheres. Reproduced with permission from [<a href="#B60-energies-17-03591" class="html-bibr">60</a>], Elsevier, 2019.</p>
Full article ">Figure 4
<p>Schematic representation of a machine learning (ML) workflow for predicting the hydrogen storage properties of materials. Reproduced with permission from [<a href="#B68-energies-17-03591" class="html-bibr">68</a>], Elsevier, 2020.</p>
Full article ">Figure 5
<p>Schematic representation of the energy landscape for hydrogen absorption and desorption in a metal hydride. Reproduced with permission from [<a href="#B77-energies-17-03591" class="html-bibr">77</a>], Springer Nature, 2023.</p>
Full article ">Figure 6
<p>Geometries of the different hydrogen trapped Li<sup>3+</sup> (types a and b) clusters optimized at B3LYP/6–311+G(d) level of theory. Reproduced with permission from [<a href="#B83-energies-17-03591" class="html-bibr">83</a>], Springer Nature, 2011.</p>
Full article ">
21 pages, 2768 KiB  
Article
System Design for Sensing in Manufacturing to Apply AI through Hierarchical Abstraction Levels
by Georgios Sopidis, Michael Haslgrübler, Behrooz Azadi, Ouijdane Guiza, Martin Schobesberger, Bernhard Anzengruber-Tanase and Alois Ferscha
Sensors 2024, 24(14), 4508; https://doi.org/10.3390/s24144508 - 12 Jul 2024
Viewed by 1070
Abstract
Activity recognition combined with artificial intelligence is a vital area of research, ranging across diverse domains, from sports and healthcare to smart homes. In the industrial domain, and the manual assembly lines, the emphasis shifts to human–machine interaction and thus to human activity [...] Read more.
Activity recognition combined with artificial intelligence is a vital area of research, ranging across diverse domains, from sports and healthcare to smart homes. In the industrial domain, and the manual assembly lines, the emphasis shifts to human–machine interaction and thus to human activity recognition (HAR) within complex operational environments. Developing models and methods that can reliably and efficiently identify human activities, traditionally just categorized as either simple or complex activities, remains a key challenge in the field. Limitations of the existing methods and approaches include their inability to consider the contextual complexities associated with the performed activities. Our approach to address this challenge is to create different levels of activity abstractions, which allow for a more nuanced comprehension of activities and define their underlying patterns. Specifically, we propose a new hierarchical taxonomy for human activity abstraction levels based on the context of the performed activities that can be used in HAR. The proposed hierarchy consists of five levels, namely atomic, micro, meso, macro, and mega. We compare this taxonomy with other approaches that divide activities into simple and complex categories as well as other similar classification schemes and provide real-world examples in different applications to demonstrate its efficacy. Regarding advanced technologies like artificial intelligence, our study aims to guide and optimize industrial assembly procedures, particularly in uncontrolled non-laboratory environments, by shaping workflows to enable structured data analysis and highlighting correlations across various levels throughout the assembly progression. In addition, it establishes effective communication and shared understanding between researchers and industry professionals while also providing them with the essential resources to facilitate the development of systems, sensors, and algorithms for custom industrial use cases that adapt to the level of abstraction. Full article
(This article belongs to the Special Issue Human-Centred Smart Manufacturing - Industry 5.0)
Show Figures

Figure 1

Figure 1
<p>This figure presents a hierarchy of an exemplary assembly process: components, units, modules, products, and post-assembly. Additionally, it demonstrates how these stages are interconnected and how activities and tasks flow inside a real assembly scenario from components to the final product. Each stage builds upon the previous one, with components being assembled into units, units into modules, modules into the final product, and finally, the product being integrated into the production line.</p>
Full article ">Figure 2
<p>The figure presents a visualization for the proposed taxonomy. At the atomic level, individual assembly activities are considered as singular tasks involving basic operations or manipulations on discrete components or tools. The micro-level aggregates multiple atomic operations into coherent sequences, representing actions within the assembly process. Larger assembly tasks are formed at the meso-level by combining multiple micro-level activities, often involving the assembly of sub-components or partial assemblies. The macro-level encompasses entire assembly processes, including stages such as the assembly of major components or modules. Finally, the mega-level represents the overall assembly process, incorporating post-assembly activities such as quality control checks, packaging, or final inspection.</p>
Full article ">Figure 3
<p>The table illustrates a simplified ATM assembly process, derived from a real industrial use case [<a href="#B76-sensors-24-04508" class="html-bibr">76</a>], showcasing activities across different assembly levels: atomic, micro, meso, macro, and mega. It serves as a comparative analysis with existing approaches for activity categorization, highlighting how each level contributes to the overall process. Specific activities are provided for clarity, offering insights into the hierarchical organization of assembly tasks. The color coding highlights differences in categorization when distinguishing tasks across levels.</p>
Full article ">Figure 4
<p>The table illustrates welding processes in car assembly, presenting the hierarchical framework of tasks, and showcasing activities across different assembly levels: atomic, micro, meso, macro, and mega. It serves as a comparative analysis with existing approaches for activity categorization, highlighting how each level contributes to the overall process and showing how individual actions aggregate into more complex tasks across the assembly line. Specific activities are provided for clarity, offering insights into the hierarchical organization of assembly tasks. The color coding highlights differences in categorization when distinguishing tasks across levels.</p>
Full article ">Figure 5
<p>This figure illustrates key characteristics across atomic, micro, meso, macro, and mega-levels of assembly activity recognition systems. Each group of related elements is color-coded, and each line represents a different category, ensuring distinctions between aspects. The figure highlights variations that are important in the overall design of an AI system, such as sensor placement, types of sensors used, system mobility, sampling rate, duration of experiments, frequency of actions, preprocessing techniques, models employed for activity recognition, window size for data processing, and feedback mechanisms. Associated recommendations are provided for each category and level to serve as a starting point for the development of AI models under the “Models to Use” category, which is related to industrial assembly.</p>
Full article ">
Back to TopTop