[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,212)

Search Parameters:
Keywords = polymer network

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
25 pages, 16876 KiB  
Article
Optimization of 3D Printing Parameters of High Viscosity PEEK/30GF Composites
by Dmitry Yu. Stepanov, Yuri V. Dontsov, Sergey V. Panin, Dmitry G. Buslovich, Vladislav O. Alexenko, Svetlana A. Bochkareva, Andrey V. Batranin and Pavel V. Kosmachev
Polymers 2024, 16(18), 2601; https://doi.org/10.3390/polym16182601 (registering DOI) - 14 Sep 2024
Viewed by 220
Abstract
The aim of this study was to optimize a set of technological parameters (travel speed, extruder temperature, and extrusion rate) for 3D printing with a PEEK-based composite reinforced with 30 wt.% glass fibers (GFs). For this purpose, both Taguchi and finite element methods [...] Read more.
The aim of this study was to optimize a set of technological parameters (travel speed, extruder temperature, and extrusion rate) for 3D printing with a PEEK-based composite reinforced with 30 wt.% glass fibers (GFs). For this purpose, both Taguchi and finite element methods (FEM) were utilized. The artificial neural networks (ANNs) were implemented for computer simulation of full-scale experiments. Computed tomography of the additively manufactured (AM) samples showed that the optimal 3D printing parameters were the extruder temperature of 460 °C, the travel speed of 20 mm/min, and the extrusion rate of 4 rpm (the microextruder screw rotation speed). These values correlated well with those obtained by computer simulation using the ANNs. In such cases, the homogeneous micro- and macro-structures were formed with minimal sample distortions and porosity levels within 10 vol.% of both structures. The most likely reason for porosity was the expansion of the molten polymer when it had been squeezed out from the microextruder nozzle. It was concluded that the mechanical properties of such samples can be improved both by changing the 3D printing strategy to ensure the preferential orientation of GFs along the building direction and by reducing porosity via post-printing treatment or ultrasonic compaction. Full article
(This article belongs to the Special Issue Additive Manufacturing of Fibre Reinforced Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>The S/N ratios for different levels of the technological parameters: (<b>a</b>) tensile strength; (<b>b</b>) elastic modulus; (<b>c</b>) elongation at break.</p>
Full article ">Figure 2
<p>The SEM micrographs of the PEEK/30GF composites additively manufactured using the modes presented in <a href="#polymers-16-02601-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>The 3D printing modes of the laboratory experiments in the space of the (input) parameters.</p>
Full article ">Figure 4
<p>The dependences of the mechanical properties of the samples of the PEEK/30 GF composite on the 3D printing parameters (<b>a</b>), as well as both dependences and histograms (<b>b</b>) after verification.</p>
Full article ">Figure 4 Cont.
<p>The dependences of the mechanical properties of the samples of the PEEK/30 GF composite on the 3D printing parameters (<b>a</b>), as well as both dependences and histograms (<b>b</b>) after verification.</p>
Full article ">Figure 5
<p>The parameters’ space and the result of checking the 3D-printing modes for compliance with the minimum acceptable property values.</p>
Full article ">Figure 6
<p>The 3D printing modes and a priori knowledge, as well as the SOP area, drawn using the RBFNN model: (<b>a</b>) spread = 0.3, goal = 0.001, the training sample size of 66 vectors; (<b>b</b>) spread = 0.3, goal = 0.01, the training sample size of 66 experimental + 54 a priori vectors.</p>
Full article ">Figure 7
<p>The experimental modes and a priori knowledge, as well as the SOP area, drawn using the FFNN model: (<b>a</b>) 4 hidden layer neurons, the sample size of 66 experimental + 132 synthesized vectors; (<b>b</b>) 6 hidden layer neurons, the sample size of 66 experimental + 54 prior + 240 synthesized vectors.</p>
Full article ">Figure 8
<p>Results of models’ verification within priory knowledge planes as a function of the size of experimental and prior vectors of the properties.</p>
Full article ">Figure 9
<p>Schematic locations of pores with the diameters of 20 µm (<b>a</b>), 100 µm (<b>b</b>), and from 20 to 100 µm (<b>c</b>) in the computational domains at the porosity of 30%.</p>
Full article ">Figure 10
<p>The elastic modulus versus porosity dependences for pores with different diameters <span class="html-italic">d</span>.</p>
Full article ">Figure 11
<p>The stress distribution surfaces over the representative volume in the presence of pores with the diameters of 20 µm (<b>a</b>), 100 µm (<b>b</b>), and from 20 to 100 µm (<b>c</b>) at a porosity of 30%.</p>
Full article ">Figure 12
<p>The three-dimensional micro-CT views of the samples, from both supporting table (<b>a</b>–<b>c</b>) and 3D-printing head (<b>d</b>–<b>f</b>) sides; mode 12 (<b>a</b>,<b>d</b>); mode 14 (<b>b</b>,<b>e</b>); mode 10 (<b>c</b>,<b>f</b>).</p>
Full article ">Figure 13
<p>The orthogonal projections of the samples near the fracture surfaces at the image (slice) sizes of 7.5–8.0 mm (<b>a</b>–<b>c</b>), 7.5–4.5 mm (<b>d</b>–<b>f</b>), and 8.0–4.5 mm (<b>g</b>–<b>i</b>). Red denotes Z- axis section; Blue denotes X-axis section; Green denotes Y-axis section.</p>
Full article ">Figure 14
<p>The comparative results of assessing the cross-sectional areas of the samples depending on the position of the height section (along the Z axis): mode 12 (sample No.19); mode 14 (sample No.28); mode 10 (sample No.30).</p>
Full article ">Figure 15
<p>Visualizations of a full tomogram (<b>a</b>) and a cut-out VOI (<b>b</b>).</p>
Full article ">Figure 16
<p>Visualizations of the areas (sections) selected for each of the modes to calculate the porosity levels for the images with sizes of 4.5–7.5 mm (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 17
<p>The orthogonal projections of an individual PEEK granule. Red denotes Z- axis section; Blue denotes X-axis section; Green denotes Y-axis section.</p>
Full article ">
45 pages, 2867 KiB  
Review
Applications of Long Short-Term Memory (LSTM) Networks in Polymeric Sciences: A Review
by Ivan Malashin, Vadim Tynchenko, Andrei Gantimurov, Vladimir Nelyub and Aleksei Borodulin
Polymers 2024, 16(18), 2607; https://doi.org/10.3390/polym16182607 (registering DOI) - 14 Sep 2024
Viewed by 181
Abstract
This review explores the application of Long Short-Term Memory (LSTM) networks, a specialized type of recurrent neural network (RNN), in the field of polymeric sciences. LSTM networks have shown notable effectiveness in modeling sequential data and predicting time-series outcomes, which are essential for [...] Read more.
This review explores the application of Long Short-Term Memory (LSTM) networks, a specialized type of recurrent neural network (RNN), in the field of polymeric sciences. LSTM networks have shown notable effectiveness in modeling sequential data and predicting time-series outcomes, which are essential for understanding complex molecular structures and dynamic processes in polymers. This review delves into the use of LSTM models for predicting polymer properties, monitoring polymerization processes, and evaluating the degradation and mechanical performance of polymers. Additionally, it addresses the challenges related to data availability and interpretability. Through various case studies and comparative analyses, the review demonstrates the effectiveness of LSTM networks in different polymer science applications. Future directions are also discussed, with an emphasis on real-time applications and the need for interdisciplinary collaboration. The goal of this review is to connect advanced machine learning (ML) techniques with polymer science, thereby promoting innovation and improving predictive capabilities in the field. Full article
(This article belongs to the Special Issue Computational and Experimental Approaches in Polymeric Materials)
19 pages, 5540 KiB  
Article
Optimizing Mechanical and Electrical Performance of SWCNTs/Fe₃O₄ Epoxy Nanocomposites: The Role of Filler Concentration and Alignment
by Zulfiqar Ali, Saba Yaqoob, Alessandro Lo Schiavo and Alberto D’Amore
Polymers 2024, 16(18), 2595; https://doi.org/10.3390/polym16182595 - 13 Sep 2024
Viewed by 210
Abstract
The demand for polymer composites with improved mechanical and electrical properties is crucial for advanced aerospace, electronics, and energy storage applications. Single-walled carbon nanotubes (SWCNTs) and iron oxide (Fe₃O₄) nanoparticles are key fillers that enhance these properties, yet challenges like orientation, uniform dispersion, [...] Read more.
The demand for polymer composites with improved mechanical and electrical properties is crucial for advanced aerospace, electronics, and energy storage applications. Single-walled carbon nanotubes (SWCNTs) and iron oxide (Fe₃O₄) nanoparticles are key fillers that enhance these properties, yet challenges like orientation, uniform dispersion, and agglomeration must be addressed to realize their full potential. This study focuses on developing SWCNTs/Fe₃O₄ epoxy composites by keeping the SWCNT concentration constant at 0.03 Vol.% and varying with Fe₃O₄ concentrations at 0.1, 0.5, and 1 Vol.% for two different configurations: randomly orientated (R-) and magnetic field-assisted horizontally aligned (A-) SWCNTs/Fe3O4 epoxy composites, and investigates the effects of filler concentration, dispersion, and magnetic alignment on the mechanical and electrical properties. The research reveals that both composite configurations achieve an optimal mechanical performance at 0.5 Vol.% Fe₃O₄, while A- SWCNTs/Fe3O4 epoxy composites outperformed at all concentrations. However, at 1 Vol.% Fe₃O₄, mechanical properties decline due to nanoparticle agglomeration, which disrupts stress distribution. In contrast, electrical conductivity peaks at 1 Vol.% Fe₃O₄, indicating that the higher density of Fe₃O₄ nanoparticles enhances the conductive network despite the mechanical losses. This study highlights the need for precise control over filler content and alignment to optimize mechanical strength and electrical conductivity in SWCNTs/Fe₃O₄ epoxy nanocomposites. Full article
(This article belongs to the Special Issue Processing, Characterization and Modeling of Polymer Nanocomposites)
19 pages, 7427 KiB  
Article
The Influence of the Structural Architecture on the Swelling Kinetics and the Network Behavior of Sodium-Alginate-Based Hydrogels Cross-Linked with Ionizing Radiation
by Ion Călina, Maria Demeter, Gabriela Crăciun, Anca Scărișoreanu and Elena Mănăilă
Gels 2024, 10(9), 588; https://doi.org/10.3390/gels10090588 - 12 Sep 2024
Viewed by 378
Abstract
The present work discusses the influence of the structural architecture of sodium alginate–co-acrylic acid–poly(ethylene) oxide hydrogels, crosslinked through electron beam (e-beam) radiation processing. The most important properties of the hydrogels were studied in detail to identify a correlation between the architecture of the [...] Read more.
The present work discusses the influence of the structural architecture of sodium alginate–co-acrylic acid–poly(ethylene) oxide hydrogels, crosslinked through electron beam (e-beam) radiation processing. The most important properties of the hydrogels were studied in detail to identify a correlation between the architecture of the hydrogels and their properties. Furthermore, the effect of sodium alginate (NaAlg) concentration, the amounts of the polymer blend, and the size of the samples on hydrogel properties were investigated. The results show that the hydrogels cross-linked (0.5% and 1% NaAlg) with 12.5 kGy exhibit improved physicochemical properties. High gel fraction levels (exceeding 83.5–93.7%) were achieved. Smaller hydrogel diameter (7 mm) contributed to a maximum swelling rate and degree of 20.440%. The hydrogel network was dependent on the hydrogels’ diameter and the amount of polymer blend used. The hydrogels best suited the first-order rate constants and exhibited a non-Fickian diffusion character with diffusion exponent values greater than 0.5. This study indicates that the cross-linked hydrogel has good properties, particularly because of its high degree of swelling and extensive stability (more than 180 h) in water. These findings show that hydrogels can be effectively applied to the purification of water contaminated with metals, dyes, or even pharmaceuticals, as well as materials with a gradual release of bioactive chemicals and water retention. Full article
(This article belongs to the Special Issue Polymeric Hydrogels for Biomedical Application)
Show Figures

Figure 1

Figure 1
<p>Representative images of hydrogels: (<b>A</b>) after e-beam irradiation at room temperature (25 °C)<span class="html-italic">;</span> (<b>B</b>); after being stored for 24 h at ambient temperature; (<b>C</b>) cut and immersed in ethanol for 24 h; (<b>D</b>) cut into discs at 15, 10, and 7 mm in diameter.</p>
Full article ">Figure 2
<p>The effect of NaAlg concentration, polymer volume, and hydrogel size on the gel fraction.</p>
Full article ">Figure 3
<p>The cross-links density of the hydrogel samples.</p>
Full article ">Figure 4
<p>Relationship between the amount of polymer solution and swelling degree of hydrogels at different sizes: (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 5
<p>The swelling degree values of the hydrogels based on size: (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 6
<p>The hydrogel samples swelled at equilibrium.</p>
Full article ">Figure 7
<p>The first-order swelling kinetics of cross-linked hydrogels: (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 8
<p>The second-order swelling kinetics of cross-linked hydrogels: (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 9
<p>The swelling kinetic curve of the cross-linked hydrogels (ln F vs. ln t): (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 10
<p>The swelling kinetic curve of the cross-linked hydrogels (F vs. t<sup>0.5</sup>): (<b>A</b>) for hydrogels with a NaAlg concentration of 0.5% and (<b>B</b>) for hydrogels with a NaAlg concentration of 1%.</p>
Full article ">Figure 11
<p>FTIR spectra of (<b>a</b>) native polymers (NaAlg, AA/ and PEO) and (<b>b</b>) I (0.5% NaAlg) and II (1% NaAlg) hydrogels with a diameter of 7 mm.</p>
Full article ">Figure 12
<p>SEM images of the (<b>a</b>) I<sub>7</sub>_20 mL, (<b>b</b>) I<sub>7</sub>_40 mL, (<b>c</b>) II<sub>7</sub>_20 mL, and (<b>d</b>) II<sub>7</sub>_40 mL hydrogels at 50× magnification. Scale bars indicate 1 mm.</p>
Full article ">
10 pages, 2127 KiB  
Article
Polymer Coating Enabled Carrier Modulation for Single-Walled Carbon Nanotube Network Inverters and Antiambipolar Transistors
by Zhao Li, Jenner H. L. Ngai and Jianfu Ding
Nanomaterials 2024, 14(18), 1477; https://doi.org/10.3390/nano14181477 - 11 Sep 2024
Viewed by 262
Abstract
The control of the performance of single-walled carbon nanotube (SWCNT) random network-based transistors is of critical importance for their applications in electronic devices, such as complementary metal oxide semiconducting (CMOS)-based logics. In ambient conditions, SWCNTs are heavily p-doped by the H2O/O [...] Read more.
The control of the performance of single-walled carbon nanotube (SWCNT) random network-based transistors is of critical importance for their applications in electronic devices, such as complementary metal oxide semiconducting (CMOS)-based logics. In ambient conditions, SWCNTs are heavily p-doped by the H2O/O2 redox couple, and most doping processes have to counteract this effect, which usually leads to broadened hysteresis and poor stability. In this work, we coated an SWCNT network with various common polymers and compared their thin-film transistors’ (TFTs’) performance in a nitrogen-filled glove box. It was found that all polymer coatings will decrease the hysteresis of these transistors due to the partial removal of charge trapping sites and also provide the stable control of the doping level of the SWCNT network. Counter-intuitively, polymers with electron-withdrawing functional groups lead to a dramatically enhanced n-branch in their transfer curve. Specifically, SWCNT TFTs with poly (vinylidene fluoride) coating show an n-type mobility up to 61 cm2/Vs, with a decent on/off ratio and small hysteresis. The inverters constructed by connecting two ambipolar TFTs demonstrate high gain but with certain voltage loss. P-type or n-type doping from polymer coating layers could suppress unnecessary n- or p-branches, shift the threshold voltage and optimize the performance of these inverters to realize rail-to-rail switching. Similar devices also demonstrate interesting antiambipolar performance with tunable on and off voltage when tested in a different configuration. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of SWCNT random network on SiO<sub>2</sub> substrate at (<b>a</b>) low and (<b>b</b>) high magnification; (<b>c</b>) schematic illustration of TFT device structure.</p>
Full article ">Figure 2
<p>Representative transfer curves of SWCNT random network TFTs: (<b>a</b>) Ctrl, (<b>b</b>) coated with PS and PVC, (<b>c</b>) coated with PMMA or Formvar, (<b>d</b>) coated with PVdF or PAN; (<b>e</b>) extracted threshold voltage for n-branch (V<sub>Tn</sub>) and corresponding electron mobility (µ<sub>e</sub>) vs. Hommett substituent constant of structure similar functional groups within polymer; (<b>f</b>) comparison of mobility and hysteresis (normalized by V<sub>G</sub> sweeping range) from PVdF-coated n-type CNT TFTs in this work with other reported random network CNT TFTs [<a href="#B6-nanomaterials-14-01477" class="html-bibr">6</a>,<a href="#B10-nanomaterials-14-01477" class="html-bibr">10</a>,<a href="#B20-nanomaterials-14-01477" class="html-bibr">20</a>,<a href="#B21-nanomaterials-14-01477" class="html-bibr">21</a>,<a href="#B22-nanomaterials-14-01477" class="html-bibr">22</a>,<a href="#B23-nanomaterials-14-01477" class="html-bibr">23</a>,<a href="#B24-nanomaterials-14-01477" class="html-bibr">24</a>,<a href="#B25-nanomaterials-14-01477" class="html-bibr">25</a>,<a href="#B26-nanomaterials-14-01477" class="html-bibr">26</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Representative transfer curves from blends of Formvar and PAN-covered SWCNT network TFTs. (<b>b</b>) Threshold voltage and electron mobility vs. weight percentage of PAN within Formvar. (<b>c</b>) P-type transfer curves from PAA or photoresistor S1813-covered SWCNT network TFTs.</p>
Full article ">Figure 4
<p>(<b>a</b>) Voltage transfer characteristics of inverter by connecting two ambipolar Ctrl TFTs; inset is circuit diagram; (<b>b</b>) gain at V<sub>DD</sub> of 14V; (<b>c</b>) hysteresis and gain vs. V<sub>DD</sub>; (<b>d</b>) output characteristics of representative Ctrl TFT.</p>
Full article ">Figure 5
<p>(<b>a</b>) Output characteristics of p-type and n-type polymer-coated CNT TFTs; (<b>b</b>) voltage transfer characteristics of inverter by connecting p/n-type TFTs, with inserted circuit diagram; (<b>c</b>) gain at various V<sub>DD</sub>; (<b>d</b>) input and output waveforms of inverter operated at 5 V.</p>
Full article ">Figure 6
<p>(<b>a</b>) Drain current—V<sub>G</sub> curves of the antiambipolar transistor by connecting two SWCNT TFTs with different polymer coating layers; the circuit diagram is shown in the inset; (<b>b</b>) a 3D plot of the drain current depending on both drain and gate bias voltage; (<b>c</b>) normalized drain current—V<sub>G</sub> curves of three antiambipolar transistors with finetuned doping levels for each TFT, with the coating polymers for each AAT shown on the left; (<b>d</b>) drain current—V<sub>G</sub> curves of a double antiambipolar transistor by connecting three SWCNT TFTs with different doping levels.</p>
Full article ">
15 pages, 7504 KiB  
Article
The Stability of UV-Defluorination-Driven Crosslinked Carbon Nanotubes: A Raman Study
by Yunxiang Gao, Mohammad Tarequl Islam, Promise Uzoamaka Otuokere, Merlyn Pulikkathara and Yuemin Liu
Nanomaterials 2024, 14(17), 1464; https://doi.org/10.3390/nano14171464 - 9 Sep 2024
Viewed by 350
Abstract
Carbon nanotubes (CNTs) are often regarded as semi-rigid, all-carbon polymers. However, unlike conventional polymers that can form 3D networks such as hydrogels or elastomers through crosslinking in solution, CNTs have long been considered non-crosslinkable under mild conditions. This perception changed with our recent [...] Read more.
Carbon nanotubes (CNTs) are often regarded as semi-rigid, all-carbon polymers. However, unlike conventional polymers that can form 3D networks such as hydrogels or elastomers through crosslinking in solution, CNTs have long been considered non-crosslinkable under mild conditions. This perception changed with our recent discovery of UV-defluorination-driven direct crosslinking of CNTs in solution. In this study, we further investigate the thermal stability of UV-defluorination-driven crosslinked CNTs, revealing that they are metastable and decompose more readily than either pristine or fluorinated CNTs under Raman laser irradiation. Using Raman spectroscopy under controlled laser power, we examined both single-walled and multi-walled fluorinated CNTs. The results demonstrate that UV-defluorinated CNTs exhibit reduced thermal stability compared to their pristine or untreated fluorinated counterparts. This instability is attributed to the strain on the intertube crosslinking bonds resulting from the curved carbon lattice of the linked CNTs. The metallic CNTs in the crosslinked CNT networks decompose and revert to their pristine state more readily than the semiconducting ones. The inherent instability of crosslinked CNTs leads to combustion at temperatures approximately 100 °C lower than those required for non-crosslinked fluorinated CNTs. This property positions crosslinked CNTs as promising candidates for applications where mechanically robust, lightweight materials are needed, along with feasible post-use removal options. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Generation and recombination of free radicals in F-SWNT bundles to form poly (CNTs), as described in our previous work [<a href="#B23-nanomaterials-14-01464" class="html-bibr">23</a>]. (<b>B</b>) Thermal stability of crosslinked CNTs studied in this work.</p>
Full article ">Figure 2
<p>(<b>A</b>) TEM images of untreated F-SWNTs and (<b>B</b>) UV-crosslinked F-SWNTs. (<b>C</b>) Raman spectra of untreated, UV-defluorinated, and hydrazine-defluorinated F-SWNTs, recorded using a 532 nm laser wavelength with a power density of 2.9 kW/mm<sup>2</sup>.</p>
Full article ">Figure 3
<p>Raman spectra of (<b>A</b>) untreated F-SWNTs and (<b>B</b>) UV-defluorinated crosslinked F-SWNTs. (<b>C</b>) D/G ratio of untreated F-SWNTs and UV-DeF-SWNTs as a function of the fraction of full Raman laser power density. The Raman laser wavelength is 532 nm, with a full laser power density of 2.9 kW/mm<sup>2</sup>. Insets in (<b>A</b>,<b>B</b>) show amplified RBM bands.</p>
Full article ">Figure 4
<p>Raman spectra of (<b>A</b>) untreated F-SWNTs and (<b>B</b>) crosslinked F-SWNTs via defluorination, using a 633 nm laser at varying power densities. (<b>C</b>) D/G ratio of untreated and crosslinked F-SWNTs as a function of the fraction of full laser power density. (<b>D</b>) Changes in RBM peak intensity in response to increasing laser power density. Full power density: 4.2 kW/mm<sup>2</sup>.</p>
Full article ">Figure 5
<p>Microscope images of the Raman laser-irradiated sample areas: (<b>A1</b>,<b>A2</b>) pristine SWNTs, (<b>B1</b>,<b>B2</b>) untreated F-SWNTs, and (<b>C1</b>,<b>C2</b>) crosslinked UV-DeF-SWNTs irradiated at 1/10 (Column 1) and full (Column 2) power density, respectively. Full power density: 2.9 kW/mm<sup>2</sup>.</p>
Full article ">Figure 6
<p>Raman spectra of (<b>A</b>) untreated F-MWNTs and (<b>B</b>) UV-DeF-MWNTs. Laser wavelength: 532 nm, with a full power density of 2.9 kW/mm<sup>2</sup>.</p>
Full article ">Figure 7
<p>Microscope images of various MWNTs irradiated with a 532 nm Raman laser at 1/10 full power density (Left, Column 1, <b>A1</b>,<b>B1</b>,<b>C1</b>) and full power density (Right, Column 2, <b>A2</b>,<b>B2</b>,<b>C2</b>). (<b>A1</b>,<b>A2</b>) Pristine MWNTs, (<b>B1</b>,<b>B2</b>) Untreated F-MWNTs, and (<b>C1</b>,<b>C2</b>) UV-defluorinated MWNTs (UV-DeF-MWNTs).</p>
Full article ">Figure 8
<p>(<b>A</b>) TGA curves of F-MWNTs and UV-DeF-MWNTs. (<b>B</b>) Enlarged view of the ignition and combustion region featured in (<b>A</b>).</p>
Full article ">
16 pages, 5004 KiB  
Article
Research on CdSe/ZnS Quantum Dots-Doped Polymer Fibers and Their Gain Characteristics
by Xuefeng Peng, Zhijian Wu and Yang Ding
Nanomaterials 2024, 14(17), 1463; https://doi.org/10.3390/nano14171463 - 9 Sep 2024
Viewed by 286
Abstract
Polymer fibers are considered ideal transmission media for all-optical networks, but their high intrinsic loss significantly limits their practical use. Quantum dot-doped polymer fiber amplifiers are emerging as a promising solution to this issue and are becoming a significant focus of research in [...] Read more.
Polymer fibers are considered ideal transmission media for all-optical networks, but their high intrinsic loss significantly limits their practical use. Quantum dot-doped polymer fiber amplifiers are emerging as a promising solution to this issue and are becoming a significant focus of research in both academia and industry. Based on the properties of CdSe/ZnS quantum dots and PMMA material, this study experimentally explores three fabrication methods for CdSe/ZnS quantum dots-doped PMMA fibers: hollow fiber filling, melt-drawing, and melt extrusion. The advantages and disadvantages of each method and key issues in fiber fabrication are analyzed. Utilizing the CdSe/ZnS quantum dots-doped PMMA fibers that were fabricated, we theoretically analyzed the key factors affecting gain performance, including fiber length, quantum dots doping concentration, and signal light intensity. Under the conditions of 1.5 W power and 445 nm laser pumping, a maximum on-off gain of 16.2 dB was experimentally achieved at 635 nm. Additionally, using a white light LED as the signal source, a broadband on-off gain with a bandwidth exceeding 70 nm and a maximum gain of 12.4 dB was observed in the 580–650 nm range. This research will contribute to the development of quantum dots-doped fiber devices and broadband optical communication technology, providing more efficient solutions for future optical communication networks. Full article
(This article belongs to the Special Issue Innovations in Nano-Based Optoelectronic Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Emission and absorption spectra and TEM image of CdSe/ZnS QDs; (<b>b</b>) the corresponding average QD diameter.</p>
Full article ">Figure 2
<p>Schematic diagram of the experimental process for fabricating CdSe/ZnS QDs-doped PMMA fibers using the hollow fiber filling method.</p>
Full article ">Figure 3
<p>Schematic diagram of the melt-drawing method for fabricating QDs-doped POFs.</p>
Full article ">Figure 4
<p>Schematic diagram of the melt extrusion method for fabricating QDs-doped POFs.</p>
Full article ">Figure 5
<p>Schematic diagram of the optical amplification testing setup for CdSe/ZnS QDs-doped PMMA fibers.</p>
Full article ">Figure 6
<p>Energy levels of the CdSe/ZnS QDs.</p>
Full article ">Figure 7
<p>(<b>a</b>) CdSe/ZnS QDs-doped PMMA fiber fabricated using the hollow fiber filling method; (<b>b</b>) bubbles in the QDs-doped fiber.</p>
Full article ">Figure 8
<p>Cross-sectional image of QDs-doped fibers fabricated using the melt-drawing method: (<b>a</b>) there are voids within the optical fiber; (<b>b</b>) the POF exhibits an asymmetric structure.</p>
Full article ">Figure 9
<p>CdSe/ZnS QDs-doped PMMA fiber fabricated using the melt extrusion method: (<b>a</b>) microscopic image of the fiber end face, (<b>b</b>) a single long fiber strand, (<b>c</b>) QDs-doped PMMA fiber under 445 nm laser pump, and (<b>d</b>) a section of the fabricated QDs-doped PMMA fiber jumper.</p>
Full article ">Figure 10
<p>Maximum gain (<b>a</b>), optimal doped fiber length (<b>b</b>), and gain spectrum (<b>c</b>) as a function of quantum dot-doped polymer fiber length.</p>
Full article ">Figure 11
<p>Relationship between gain, doped fiber length, and QDs doping concentration.</p>
Full article ">Figure 12
<p>Variation of gain with signal light intensity.</p>
Full article ">Figure 13
<p>Gain spectra of CdSe/ZnS QDs-doped PMMA fiber amplifier at different pump energy densities.</p>
Full article ">Figure 14
<p>Experimental on-off gain as a function of pump power.</p>
Full article ">Figure 15
<p>Variation of on-off gain with pump power at different signal light intensities.</p>
Full article ">Figure 16
<p>Variation of on-off gain spectra with pump power, inset showing the LED light spectrum.</p>
Full article ">
13 pages, 7271 KiB  
Article
In Situ Analysis of Binder Degradation during Catalyst-Accelerated Stress Test of Polymer Electrolyte Membrane Fuel Cells
by Donggeun Yoo, Sujung Park, Sohyeong Oh, Minsoo P. Kim and Kwonpil Park
Materials 2024, 17(17), 4425; https://doi.org/10.3390/ma17174425 - 9 Sep 2024
Viewed by 327
Abstract
High-oxygen-permeability ionomers (HOPIs) are being actively developed to enhance the performance and durability of high-power polymer electrolyte membrane fuel cells (PEMFCs). While methods for evaluating binder performance are well-established, techniques for assessing binder durability and measuring its degradation in situ during the AST [...] Read more.
High-oxygen-permeability ionomers (HOPIs) are being actively developed to enhance the performance and durability of high-power polymer electrolyte membrane fuel cells (PEMFCs). While methods for evaluating binder performance are well-established, techniques for assessing binder durability and measuring its degradation in situ during the AST process remain limited. This study examines the distribution of relaxation times (DRT) and Warburg-like response (WLR) methods as in situ analysis techniques during the catalyst-accelerated stress test (AST) process. We conducted catalyst-ASTs (0.6–0.95 V cycling) for 20,000 cycles, monitoring changes using cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and linear sweep voltammetry (LSV). Contrary to expectations, during the catalyst-AST, the ion transport resistance of the binder decreased, indicating no binder degradation. Scanning electron microscopy/energy dispersive spectrometer (SEM/EDS) analysis revealed that the degradation rate of the catalyst and the support was relatively higher than that of the binder, leading to a reduction in catalyst layer thickness and improved binder network formation. By applying the DRT method during the catalyst-AST process, we were able to measure the increase in oxygen reduction reaction (ORR) resistance and the decrease in proton transport resistance in situ. This allowed for the real-time detection of the reduction in catalyst layer thickness and improvements in ionomer networks due to catalyst and support degradation. These findings provide new insights into the complex interplay between catalyst degradation and binder performance, contributing to the development of more durable PEMFC components. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) CV changes during AST. (<b>b</b>) ECSA and (<b>c</b>) DLC changes for AST cycles.</p>
Full article ">Figure 2
<p>(<b>a</b>) Impedance changes during AST. (<b>b</b>) CTR and (<b>c</b>) HFR changes for AST cycles.</p>
Full article ">Figure 3
<p>(<b>a</b>) LSV changes during AST. (<b>b</b>) HCCD and (<b>c</b>) short resistance changes for AST cycles.</p>
Full article ">Figure 4
<p>(<b>a</b>) I–V curve changes during AST. (<b>b</b>) OCV and (<b>c</b>) current density @ 0.6 V changes for AST cycles.</p>
Full article ">Figure 5
<p>(<b>a</b>) DRT changes during AST. (<b>b</b>) Magnification of proton transfer resistance peak. (<b>c</b>) ORR charge transfer resistance and proton transfer resistance in CCL changes for AST cycles.</p>
Full article ">Figure 6
<p>(<b>a</b>) WLR impedance changes during AST. (<b>b</b>) WLR changes for AST cycles.</p>
Full article ">Figure 7
<p>Cross-sectional SEM images of MEA after different AST cycles: (<b>a</b>) pristine, (<b>b</b>) after 5000 AST cycles, (<b>c</b>) after 10,000 AST cycles, and (<b>d</b>) after 20,000 AST cycles.</p>
Full article ">Figure 8
<p>Cross-sectional SEM images of cathode catalyst layers: (<b>a</b>) pristine and (<b>b</b>) after 20,000 AST cycles. EDS mapping of the rectangular area within the cathode catalyst layer in the SEM image: (<b>c</b>) pristine and (<b>c</b>) after 20,000 AST cycles. Magnified images of the rectangular areas in (<b>c</b>) and (<b>d</b>) are shown in (<b>e</b>) and (<b>f</b>), respectively.</p>
Full article ">
21 pages, 2290 KiB  
Article
Red-Billed Blue Magpie Optimizer for Electrical Characterization of Fuel Cells with Prioritizing Estimated Parameters
by Attia A. El-Fergany and Ahmed M. Agwa
Technologies 2024, 12(9), 156; https://doi.org/10.3390/technologies12090156 - 8 Sep 2024
Viewed by 529
Abstract
The red-billed blue magpie optimizer (RBMO) is employed in this research study to address parameter extraction in polymer exchange membrane fuel cells (PEMFCs), along with three recently implemented optimizers. The sum of squared deviations (SSD) between the simulated and measured stack voltages defines [...] Read more.
The red-billed blue magpie optimizer (RBMO) is employed in this research study to address parameter extraction in polymer exchange membrane fuel cells (PEMFCs), along with three recently implemented optimizers. The sum of squared deviations (SSD) between the simulated and measured stack voltages defines the fitness function of the optimization problem under investigation subject to a set of working constraints. Three distinct PEMFCs stacks models—the Ballard Mark, Temasek 1 kW, and Horizon H-12 units—are used to illustrate the applied RBMO’s feasibility in solving this challenge in comparison to other recent algorithms. The highest percentages of biased voltage per reading for the Ballard Mark V, Temasek 1 kW, and Horizon H-12 are, respectively, +0.65%, +0.20%, and −0.14%, which are negligible errors. The primary characteristics of PEMFC stacks under changing reactant pressures and cell temperatures are used to evaluate the precision of the cropped optimized parameters. In the final phase of this endeavor, the sensitivity of the cropped parameters to the PEMFCs model’s performance is investigated using two machine learning techniques, namely, artificial neural network and Gaussian process regression models. The simulation results demonstrate that the RBMO approach extracts the PEMFCs’ appropriate parameters with high precision. Full article
(This article belongs to the Collection Electrical Technologies)
Show Figures

Figure 1

Figure 1
<p>Procedures of the RBMO framework.</p>
Full article ">Figure 2
<p>Convergence patterns of all studied test cases.</p>
Full article ">Figure 3
<p>The principal performance of Ballard Mark V.</p>
Full article ">Figure 4
<p>Principal performance of Temasek 1 kW.</p>
Full article ">Figure 5
<p>Principal performance of Horizon H-12 unit.</p>
Full article ">Figure 6
<p>Percentage of biased voltage.</p>
Full article ">Figure 6 Cont.
<p>Percentage of biased voltage.</p>
Full article ">
5 pages, 4947 KiB  
Proceeding Paper
Assessing Viscoelastic Parameters of Polymer Pipes via Transient Signals and Artificial Neural Networks
by Mostafa Rahmanshahi, Huan-Feng Duan, Alireza Keramat, Nasim Vafaei Rad and Hossein Azizi Nadian
Eng. Proc. 2024, 69(1), 74; https://doi.org/10.3390/engproc2024069074 - 6 Sep 2024
Viewed by 118
Abstract
This study presents a soft-computing-based method for determining polymer pipelines’ creep function parameters (CFPs) and pressure wave speeds (PWSs) through transient flow analysis. To this end, first, a numerical model for transient flow in polymer pipes was developed in the time domain. Then, [...] Read more.
This study presents a soft-computing-based method for determining polymer pipelines’ creep function parameters (CFPs) and pressure wave speeds (PWSs) through transient flow analysis. To this end, first, a numerical model for transient flow in polymer pipes was developed in the time domain. Then, by considering a pipeline with a specific geometry, 2000 transient flow signals were generated for different CFPs and PWSs. The amplitudes obtained by transforming the time-domain pressure signals to the frequency domain using the fast Fourier transform algorithm are the inputs for an artificial neural network model. The results showed that the proposed approach accurately estimated the creep function of the polymer pipes. Full article
Show Figures

Figure 1

Figure 1
<p>Proposed ANN-based CFP and PWS prediction.</p>
Full article ">Figure 2
<p>Training, validation, and testing residuals and absolute errors for (<b>a</b>) <span class="html-italic">j<sub>1</sub></span>; (<b>b</b>) <span class="html-italic">j<sub>2</sub></span>; (<b>c</b>) <span class="html-italic">j<sub>3</sub></span><b>;</b> (<b>d</b>) <span class="html-italic">a</span>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Time; (<b>b</b>) frequency domain pressure signals of Test #1 and Test #2.</p>
Full article ">Figure 4
<p>Comparison between original data and ANN results for (<b>a</b>) Test #1; (<b>b</b>) Test #2.</p>
Full article ">
17 pages, 5470 KiB  
Article
Microfiltration Membrane Pore Functionalization with Primary and Quaternary Amines for PFAS Remediation: Capture, Regeneration, and Reuse
by Sam Thompson, Angela M. Gutierrez, Jennifer Bukowski and Dibakar Bhattacharyya
Molecules 2024, 29(17), 4229; https://doi.org/10.3390/molecules29174229 - 6 Sep 2024
Viewed by 381
Abstract
The widespread production and use of multi-fluorinated carbon-based substances for a variety of purposes has contributed to the contamination of the global water supply in recent decades. Conventional wastewater treatment can reduce contaminants to acceptable levels, but the concentrated retentate stream is still [...] Read more.
The widespread production and use of multi-fluorinated carbon-based substances for a variety of purposes has contributed to the contamination of the global water supply in recent decades. Conventional wastewater treatment can reduce contaminants to acceptable levels, but the concentrated retentate stream is still a burden to the environment. A selective anion-exchange membrane capable of capture and controlled release could further concentrate necessary contaminants, making their eventual degradation or long-term storage easier. To this end, commercial microfiltration membranes were modified using pore functionalization to incorporate an anion-exchange moiety within the membrane matrix. This functionalization was performed with primary and quaternary amine-containing polymer networks ranging from weak to strong basic residues. Membrane loading ranged from 0.22 to 0.85 mmol/g membrane and 0.97 to 3.4 mmol/g membrane for quaternary and primary functionalization, respectively. Modified membranes exhibited a range of water permeances within approximately 45–131 LMH/bar. The removal of PFASs from aqueous streams was analyzed for both “long-chain” and “short-chain” analytes, perfluorooctanoic acid and perfluorobutyric acid, respectively. Synthesized membranes demonstrated as high as 90% rejection of perfluorooctanoic acid and 50–80% rejection of perfluorobutyric acid after 30% permeate recovery. Regenerated membranes maintained the capture performance for three cycles of continuous operation. The efficiency of capture and reuse can be improved through the consideration of charge density, water flux, and influent contaminant concentration. This process is not limited by the substrate and, thus, is able to be implemented on other platforms. This research advances a versatile membrane platform for environmentally relevant applications that seek to help increase the global availability of safe drinking water. Full article
(This article belongs to the Section Green Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic of functionalization monomers used for membrane synthesis, targeted contaminant capture, and subsequent regeneration. Not to scale.</p>
Full article ">Figure 2
<p>Graph of transmittance intensity attained using Fourier-transform infrared spectroscopy of isolated polymer hydrogel and functionalized membranes. Coordination of transmittance peaks corresponds to presence of chemical bonds.</p>
Full article ">Figure 3
<p>Scanning electron microscopy of functionalized membrane surface. (<b>a</b>) Pristine PVDF membrane, scale bar 500 nm; (<b>b</b>) larger view of PVDF membrane functionalized with DMAPA-Q, scale bar 5 µm; (<b>c</b>) hydrophilized PES membrane functionalized with DMAPA-Q, scale bar 5 µm; (<b>d</b>) PVDF membrane functionalized with PAH, scale bar 5 µm.</p>
Full article ">Figure 4
<p>Effect of feed pH on water permeance. Permeation operated in dead-end batch conditions. pH adjusted with hydrochloric acid or sodium hydroxide. Error bars represent +/− one standard deviation in series of permeance measurements, <span class="html-italic">n</span> = 3.</p>
Full article ">Figure 5
<p>PFOA/PFBA capture by primary and quaternary amine-functionalized PVDF650 membrane and pristine PVDF650 blank membrane using dead-end batch filtration. Capture efficiency is defined as ((feed concentration—permeate concentration)/feed concentration) × 100%. Permeate represents 30% water recovery. (<b>a</b>) PFOA capture efficiency; (<b>b</b>) PFBA capture efficiency; (<b>c</b>) PFOA capture efficiency during processing; dotted 45° line represents capture efficiency of 100%. Error bars represent +/− one standard deviation using technical triplicates; for PVDF blank capture, <span class="html-italic">n</span> = 1.</p>
Full article ">Figure 6
<p>Regeneration efficiency of functionalized anion-exchange membranes, defined as percentage of captured PFASs released in approximately 25% regeneration volume. DMAPAQ-functionalized PVDF and unfunctionalized PVDF regeneration stream employed a 50:50 methanol and water mixture, while PAH-functionalized PVDF regeneration stream used water with a pH of 10.5. (<b>a</b>) PFOA concentration in permeate using regeneration stream; capture experiment employed 120 ppb PFOA; (<b>b</b>) PFBA concentration in permeate using regeneration stream; capture experiment employed 140 ppb PFBA. All experiments were performed at 4.8 bar (LMH/bar for functionalized membrane = 50–120); V in captions indicates volume, <span class="html-italic">n</span> = 1. PFOA analytical error: +/−5.5%; PFBA analytical error: +/−6.5%.</p>
Full article ">Figure 7
<p>Capture and controlled release of (<b>a</b>) 120 ppb PFOA solution and (<b>b</b>) 140 ppb PFBA solution in functionalized and pristine membranes. Filled-in symbols indicate concentration of PFASs detected in permeate. Open symbols indicate concentration of PFAS detected in the regeneration stream. Green region indicates permeate from ~130 ppb PFOA/PFBA mixture, red region indicates permeate from regeneration stream. DMAPA-Q and PAH membranes were regenerated with 50:50 methanol/water mix and water adjusted to pH 10.5, respectively. All permeations were performed at 4.8 bar under neutral conditions, <span class="html-italic">n</span> = 1. * Denotes measurement below instrumentation limit of quantification.</p>
Full article ">
18 pages, 4453 KiB  
Article
Electrospun PVP Fibers as Carriers of Ca2+ Ions to Improve the Osteoinductivity of Titanium-Based Dental Implants
by Janina Roknić, Ines Despotović, Jozefina Katić and Željka Petrović
Molecules 2024, 29(17), 4181; https://doi.org/10.3390/molecules29174181 - 3 Sep 2024
Viewed by 654
Abstract
Although titanium and its alloys are widely used as dental implants, they cannot induce the formation of new bone around the implant, which is a basis for the functional integrity and long-term stability of implants. This study focused on the functionalization of the [...] Read more.
Although titanium and its alloys are widely used as dental implants, they cannot induce the formation of new bone around the implant, which is a basis for the functional integrity and long-term stability of implants. This study focused on the functionalization of the titanium/titanium oxide surface as the gold standard for dental implants, with electrospun composite fibers consisting of polyvinylpyrrolidone and Ca2+ ions. Polymer fibers as carriers of Ca2+ ions should gradually dissolve, releasing Ca2+ ions into the environment of the implant when it is immersed in a model electrolyte of artificial saliva. Scanning electron microscopy, energy dispersive X-ray spectroscopy and attenuated total reflectance Fourier transform infrared spectroscopy confirmed the successful formation of a porous network of composite fibers on the titanium/titanium oxide surface. The mechanism of the formation of the composite fibers was investigated in detail by quantum chemical calculations at the density functional theory level based on the simulation of possible molecular interactions between Ca2+ ions, polymer fibers and titanium substrate. During the 7-day immersion of the functionalized titanium in artificial saliva, the processes on the titanium/titanium oxide/composite fibers/artificial saliva interface were monitored by electrochemical impedance spectroscopy. It can be concluded from all the results that the composite fibers formed on titanium have application potential for the development of osteoinductive and thus more biocompatible dental implants. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The ATR-FTIR spectra of (<b>a</b>) CaCl<sub>2</sub> salt, PVP and (PVP+Ca<sup>2+</sup>) fibers; (<b>b</b>) freshly prepared Ti, Ti modified with thermally prepared oxide film (Ti/Ti oxide) and Ti modified with the composite fibers [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers].</p>
Full article ">Figure 2
<p>The SEM images and corresponding EDS spectra of (<b>a</b>,<b>b</b>) (PVP+Ca<sup>2+</sup>) fibers; (<b>c</b>,<b>d</b>) Ti modified with thermally prepared oxide film (Ti/Ti oxide) and (<b>e</b>,<b>f</b>) Ti modified with the composite fibers [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers].</p>
Full article ">Figure 3
<p>The optimized structures of (<b>a</b>) PVP-Ca-I, in which the PVP-Ca bond is established via a nitrogen atom; (<b>b</b>) PVP-Ca-II, in which the PVP-Ca bond is established via oxygen atoms of the carbonyl groups. The bond distances are given in Å. The bond energies are given in kcal mol<sup>−1</sup>. O—red, C—gray, N—blue, H—white, Ca—yellow-green.</p>
Full article ">Figure 4
<p>The optimized structures of (<b>a</b>) PVP-Ca-TiO<sub>2</sub>-I, in which the Ca<sup>2+</sup> is complexed in between PVP layer and TiO<sub>2</sub> surface, (<b>b</b>) PVP-Ca-TiO<sub>2</sub>-II, in which the Ca<sup>2+</sup> is bound on the top of PVP layer. The bond distances are given in Å. The bond energies are given in kcal mol<sup>−1</sup>. O—red, C—gray, N—blue, H—white, Ca—yellow-green.</p>
Full article ">Figure 5
<p>The EIS plots in the form of (<b>a</b>) magnitude vs. log <span class="html-italic">f,</span> (<b>b</b>) phase angle vs. log <span class="html-italic">f</span> recorded on the [Ti/Ti oxide/(PVP+Ca<sup>2+</sup>) fibers] sample at open circuit potential after 1 h, 2 days and 7 days of immersion in artificial saliva, pH = 6.8.</p>
Full article ">Figure 6
<p>The SEM images and corresponding EDS spectra of (<b>a</b>,<b>b</b>) freshly electrospun (PVP+Ca<sup>2+</sup>) fibers on Al foil; (<b>c</b>,<b>d</b>) the Al foil surface with residual fibers after 7 days of immersion in artificial saliva, pH = 6.8. Fe, visible in the spectrum (<b>d</b>), originates from Al foil.</p>
Full article ">Figure 7
<p>The photographs of (<b>a</b>) the freshly abraded and degreased surface of the Ti samples; (<b>b</b>) the thermally generated oxide film on the Ti (Ti/Ti oxide); and (<b>c</b>) the (PVP+Ca<sup>2+</sup>) fibers on the Ti/Ti oxide sample.</p>
Full article ">
12 pages, 1539 KiB  
Article
Gel Diffusiophoresis of a Spherical Colloidal Particle
by Hiroyuki Ohshima
Fluids 2024, 9(9), 203; https://doi.org/10.3390/fluids9090203 - 1 Sep 2024
Viewed by 455
Abstract
A theoretical framework is established for the gel diffusiophoresis of a spherical colloidal particle moving through an uncharged dilute porous polymer gel medium when an electrolyte concentration gradient field is applied. The network of cross-linked polymer segments is treated as a porous skeleton [...] Read more.
A theoretical framework is established for the gel diffusiophoresis of a spherical colloidal particle moving through an uncharged dilute porous polymer gel medium when an electrolyte concentration gradient field is applied. The network of cross-linked polymer segments is treated as a porous skeleton containing an electrolyte solution using the Brinkman–Debye–Bueche model. We derive a general expression for the gel-diffusiophoretic mobility of a charged spherical colloidal particle. Based on this general mobility expression, we farther derive a closed-form approximate expression for the gel-diffusiophoretic mobility of a weakly charged spherical particle correct to the second order of the particle’s zeta potential. The obtained mobility expression depends on the Debye–Hückel parameter and the Brinkmann parameter. Full article
Show Figures

Figure 1

Figure 1
<p>Gel diffusiophoresis of a spherical colloidal particle with a radius <span class="html-italic">a</span> moving with a diffusiophoretic velocity <b><span class="html-italic">U</span></b> in an electrolyte concentration gradient field ∇<span class="html-italic">n</span><sup>∞</sup> or <b><span class="html-italic">α</span></b>.</p>
Full article ">Figure 2
<p>Scaled gel-diffusiophoretic mobility <span class="html-italic">U</span>* of a spherical colloidal particle with a radius <span class="html-italic">a</span> moving in an uncharged dilute porous polymer gel medium containing an aqueous electrolyte solution at 25 °C plotted as a function of the scaled zeta potential <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ζ</mi> </mrow> <mo stretchy="false">~</mo> </mover> </mrow> </semantics></math>. Calculated via Equation (53) for several values of <span class="html-italic">λa</span> at <span class="html-italic">κa</span> = 10 (<b>a</b>,<b>b</b>) and <span class="html-italic">κa</span> = 50 (<b>c</b>,<b>d</b>). Results for KCl (<span class="html-italic">m</span><sub>+</sub> = 0.176, <span class="html-italic">m</span><sub>−</sub> = 0.169, <span class="html-italic">β</span> = −0.02) are shown in (<b>a</b>,<b>c</b>), and those for NaCl (<span class="html-italic">m</span><sub>+</sub> = 0.258, <span class="html-italic">m</span><sub>−</sub> = 0.169, <span class="html-italic">β</span> = −0. 2) are shown in (<b>b</b>,<b>d</b>). The curve with <span class="html-italic">λa</span> = 0 corresponds to the scaled diffusiophoretic mobility in a free electrolyte solution.</p>
Full article ">Figure 2 Cont.
<p>Scaled gel-diffusiophoretic mobility <span class="html-italic">U</span>* of a spherical colloidal particle with a radius <span class="html-italic">a</span> moving in an uncharged dilute porous polymer gel medium containing an aqueous electrolyte solution at 25 °C plotted as a function of the scaled zeta potential <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ζ</mi> </mrow> <mo stretchy="false">~</mo> </mover> </mrow> </semantics></math>. Calculated via Equation (53) for several values of <span class="html-italic">λa</span> at <span class="html-italic">κa</span> = 10 (<b>a</b>,<b>b</b>) and <span class="html-italic">κa</span> = 50 (<b>c</b>,<b>d</b>). Results for KCl (<span class="html-italic">m</span><sub>+</sub> = 0.176, <span class="html-italic">m</span><sub>−</sub> = 0.169, <span class="html-italic">β</span> = −0.02) are shown in (<b>a</b>,<b>c</b>), and those for NaCl (<span class="html-italic">m</span><sub>+</sub> = 0.258, <span class="html-italic">m</span><sub>−</sub> = 0.169, <span class="html-italic">β</span> = −0. 2) are shown in (<b>b</b>,<b>d</b>). The curve with <span class="html-italic">λa</span> = 0 corresponds to the scaled diffusiophoretic mobility in a free electrolyte solution.</p>
Full article ">
14 pages, 10170 KiB  
Article
AgNP Composite Silicone-Based Polymer Self-Healing Antifouling Coatings
by Xingda Liu, Jiawen Sun, Jizhou Duan, Kunyan Sui, Xiaofan Zhai and Xia Zhao
Materials 2024, 17(17), 4289; https://doi.org/10.3390/ma17174289 - 30 Aug 2024
Viewed by 333
Abstract
Biofouling poses a significant challenge to the marine industry, and silicone anti-biofouling coatings have garnered extensive attention owing to their environmental friendliness and low surface energy. However, their widespread application is hindered by their low substrate adhesion and weak static antifouling capabilities. In [...] Read more.
Biofouling poses a significant challenge to the marine industry, and silicone anti-biofouling coatings have garnered extensive attention owing to their environmental friendliness and low surface energy. However, their widespread application is hindered by their low substrate adhesion and weak static antifouling capabilities. In this study, a novel silicone polymer polydimethylsiloxane (PDMS)-based poly(urea-thiourea-imine) (PDMS-PUTI) was synthesized via stepwise reactions of aminopropyl-terminated polydimethylsiloxane (APT-PDMS) with isophorone diisocyanate (IPDI), isophthalaldehyde (IPAL), and carbon disulfide (CS2). Subsequently, a nanocomposite coating (AgNPs-x/PDMS-PUTI) was prepared by adding silver nanoparticles (AgNPs) to the polymer PDMS-PUTI. The dynamic multiple hydrogen bonds formed between urea and thiourea linkages, along with dynamic imine bonds in the polymer network, endowed the coating with outstanding self-healing properties, enabling complete scratch healing within 10 min at room temperature. Moreover, uniformly dispersed AgNPs not only reduced the surface energy of the coating but also significantly enhanced its antifouling performance. The antibacterial efficiency against common marine bacteria Pseudomonas aeruginosa (P.sp) and Staphylococcus aureus (S.sp) was reduced by 97.08% and 96.71%, respectively, whilst the diatom settlement density on the coating surface was as low as approximately 59 ± 3 diatom cells/mm2. This study presents a novel approach to developing high-performance silicone antifouling coatings. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Synthesis of AgNPs-X/PDMS-PUTI. (<b>b</b>) AgNPs-X/PDMS-PUTI design diagram.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Synthesis of AgNPs-X/PDMS-PUTI. (<b>b</b>) AgNPs-X/PDMS-PUTI design diagram.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra of PDMS-PUTI.</p>
Full article ">Figure 3
<p>FTIR spectra of PDMS-PUTI.</p>
Full article ">Figure 4
<p>GPC curve of PDMS-PUTI.</p>
Full article ">Figure 5
<p>(<b>a</b>) Appearance of PDMS-PUTI and AgNPs-x/PDMS-PUTI. (<b>b</b>) The stress–strain curves at 25 °C of PDMS-PUTI and AgNPs-x/PDMS-PUTI.</p>
Full article ">Figure 6
<p>Self-healing properties of the AgNPs-x/PDMS-PUTI coatings. Micrographs images of the self-healing process of (<b>a</b>) PDMS-PUTI, (<b>b</b>) AgPNs-3/PDMS-PUTI, (<b>c</b>) AgPNs-6/PDMS-PUTI, and (<b>d</b>) AgPNs-9/PDMS-PUTI at 25 °C in the air (scratch thickness: ~120 μm, scratch width: ~7 μm, film thickness: ~0.7 mm).</p>
Full article ">Figure 7
<p>Adhesion strength of PDMS and AgNPs-x/PDMS-PUTI coatings adhered to the GFE and steel.</p>
Full article ">Figure 8
<p>SEM and EDS spectrum images (C, S, Cu, Si, Ag) of the surface of (<b>a</b>) PDMS-PUTI, (<b>b</b>) AgNPs-3/PDMS-PUTI, (<b>c</b>) AgNPs-6/PDMS-PUTI, (<b>d</b>) AgNPs-9/PDMS-PUTI.</p>
Full article ">Figure 9
<p>CLSM images of (<b>a</b>) PDMS-PUTI, (<b>b</b>) AgNPs-3/PDMS-PUTI, (<b>c</b>) AgNPs-3/PDMS-PUTI, and (<b>d</b>) AgNPs-3/PDMS-PUTI (image showing a 324 μm × 322 μm area).</p>
Full article ">Figure 10
<p>(<b>a</b>) Images of water and DIM contact angles for PDMS-PUTI, PDMS-PUTI/0.1, PDMS-PUTI/0.5, and PDMS-PUTI/1.0, (<b>b</b>) WCA of PDMS, PDMS-PUTI, PDMS-PUTI/0.1, PDMS-PUTI/0.5, and PDMS-PUTI/1.0, (<b>c</b>) SE of PDMS, PDMS-PUTI, PDMS-PUTI/0.1, PDMS-PUTI/0.5, and PDMS-PUTI/1.0.</p>
Full article ">Figure 11
<p>Removal strength of pseudobarnacles on each coating.</p>
Full article ">Figure 12
<p>Antibacterial and anti-diatom properties of AgNPs-X/PDMS-PUTI coatings. Fluorescence images of (<b>a</b>) <span class="html-italic">P.</span>sp. and (<b>b</b>) <span class="html-italic">S</span>.sp, (<b>c</b>) <span class="html-italic">N. incerta</span> adhering to PDMS, PDMS-PUTI, AgNPs-3/PDMS-PUTI, AgNPs-6/PDMS-PUTI, and AgNPs-9/PDMS-PUTI, (<b>d</b>) quantitative evaluation of <span class="html-italic">P.</span>sp. and <span class="html-italic">S.</span>sp. adhesion rates, (<b>e</b>) quantitative colonization density of <span class="html-italic">N. incerta</span> on coating surfaces.</p>
Full article ">
10 pages, 3459 KiB  
Article
Prediction of Glass Transition Temperature of Polymers Using Simple Machine Learning
by Jaka Fajar Fatriansyah, Baiq Diffa Pakarti Linuwih, Yossi Andreano, Intan Septia Sari, Andreas Federico, Muhammad Anis, Siti Norasmah Surip and Mariatti Jaafar
Polymers 2024, 16(17), 2464; https://doi.org/10.3390/polym16172464 - 29 Aug 2024
Viewed by 823
Abstract
Polymer materials have garnered significant attention due to their exceptional mechanical properties and diverse industrial applications. Understanding the glass transition temperature (Tg) of polymers is critical to prevent operational failures at specific temperatures. Traditional methods for measuring Tg, [...] Read more.
Polymer materials have garnered significant attention due to their exceptional mechanical properties and diverse industrial applications. Understanding the glass transition temperature (Tg) of polymers is critical to prevent operational failures at specific temperatures. Traditional methods for measuring Tg, such as differential scanning calorimetry (DSC) and dynamic mechanical analysis, while accurate, are often time-consuming, costly, and susceptible to inaccuracies due to random and uncertain factors. To address these limitations, the aim of the present study is to investigate the potential of Simplified Molecular Input Line Entry System (SMILES) as descriptors in simple machine learning models to predict Tg efficiently and reliably. Five models were utilized: k-nearest neighbors (KNNs), support vector regression (SVR), extreme gradient boosting (XGBoost), artificial neural network (ANN), and recurrent neural network (RNN). SMILES descriptors were converted into numerical data using either One Hot Encoding (OHE) or Natural Language Processing (NLP). The study found that SMILES inputs with fewer than 200 characters were inadequate for accurately describing compound structures, while inputs exceeding 200 characters diminished model performance due to the curse of dimensionality. The ANN model achieved the highest R2 value of 0.79; however, the XGB model, with an R2 value of 0.774, exhibited the highest stability and shorter training times compared to other models, making it the preferred choice for Tg prediction. The efficiency of the OHE method over NLP was demonstrated by faster training times across the KNN, SVR, XGB, and ANN models. Validation of new polymer data showed the XGB model’s robustness, with an average prediction deviation of 9.76 from actual Tg values. These findings underscore the importance of optimizing SMILES conversion methods and model parameters to enhance prediction reliability. Future research should focus on improving model accuracy and generalizability by incorporating additional features and advanced techniques. This study contributes to the development of efficient and reliable predictive models for polymer properties, facilitating the design and application of new polymer materials. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of the relationship between SMILES character length and model performance.</p>
Full article ">Figure 2
<p>Performance and data distribution of the KNN (<b>a</b>), SVR (<b>b</b>), XGBoost (<b>c</b>), ANN (<b>d</b>), and RNN (<b>e</b>) models trained with the optimal parameters. The blue dots represent the testing data points, while the red line denotes a condition where the true value is the same as predicted value.</p>
Full article ">Figure 2 Cont.
<p>Performance and data distribution of the KNN (<b>a</b>), SVR (<b>b</b>), XGBoost (<b>c</b>), ANN (<b>d</b>), and RNN (<b>e</b>) models trained with the optimal parameters. The blue dots represent the testing data points, while the red line denotes a condition where the true value is the same as predicted value.</p>
Full article ">Figure 3
<p>Comparison of model performance stability over 10 iterations using K-fold method.</p>
Full article ">Figure 4
<p>The <span class="html-italic">T<sub>g</sub></span> polymer data distribution.</p>
Full article ">
Back to TopTop