[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
remove_circle_outline

Search Results (3,097)

Search Parameters:
Keywords = wettability

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 11628 KiB  
Article
A Comprehensive Evaluation of Electrochemical Performance of Aluminum Hybrid Nanocomposites Reinforced with Alumina (Al2O3) and Graphene Oxide (GO)
by Muhammad Faizan Khan, Abdul Samad Mohammed and Ihsan-ul-Haq Toor
Metals 2024, 14(9), 1057; https://doi.org/10.3390/met14091057 - 16 Sep 2024
Abstract
The electrochemical performance of in-house developed, spark plasma-sintered, Aluminum metal–matrix composites (MMCs) was evaluated using different electrochemical techniques. X-ray diffraction (XRD) and Raman spectra were used to characterize the nanocomposites along with FE-SEM and EDS for morphological, structural, and elemental analysis, respectively. The [...] Read more.
The electrochemical performance of in-house developed, spark plasma-sintered, Aluminum metal–matrix composites (MMCs) was evaluated using different electrochemical techniques. X-ray diffraction (XRD) and Raman spectra were used to characterize the nanocomposites along with FE-SEM and EDS for morphological, structural, and elemental analysis, respectively. The highest charge transfer resistance (Rct), lowest corrosion current density, lowest electrochemical potential noise (EPN), and electrochemical current noise (ECN) were observed for GO-reinforced Al-MMC. The addition of honeycomb-like structure in the Al matrix assisted in blocking the diffusion of Cl or SO4−2. However, poor wettability in between Al matrix and Al2O3 reinforcement resulted in the formation of porous interface regions, leading to a degradation in the corrosion resistance of the composite. Post-corrosion surface analysis by optical profilometer indicated that, unlike its counterparts, the lowest surface roughness (Ra) was provided by GO-reinforced MMC. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the synthesis of nanocomposite and hybrid nanocomposite from raw nanomaterials through spark plasma sintering.</p>
Full article ">Figure 2
<p>(<b>a</b>) Relative and theoretical densities, (<b>b</b>) Hardness of pure Al, Al-0.25% GO, Al-10 vol% Al<sub>2</sub>O<sub>3</sub>, and Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO.</p>
Full article ">Figure 3
<p>(<b>a</b>) X-ray diffraction (XRD), (<b>b</b>) Raman spectra of metallic Al, Al-0.25% GO, Al-10% Al<sub>2</sub>O<sub>3</sub>, and Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO (hybrid-) nanocomposites.</p>
Full article ">Figure 4
<p>SEM for the as-received powders of (<b>a</b>) Al, (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>) GO.</p>
Full article ">Figure 5
<p>The 3D-SEM images of the Al based nanocomposites after SPS. (<b>a</b>) Pure Al, (<b>b</b>) Al-0.25% GO, (<b>c</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>. (<b>d</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO.</p>
Full article ">Figure 6
<p>EDS of the Al-based nanocomposites after SPS. (<b>a</b>) Pure Al, (<b>b</b>) Al-0.25% GO, (<b>c</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>. (<b>d</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO.</p>
Full article ">Figure 7
<p>Contact angles measured on (<b>a</b>) Al, (<b>b</b>) Al-0.25% GO, (<b>c</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>, (<b>d</b>) Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO.</p>
Full article ">Figure 8
<p>Electrochemical equivalent circuits used for fitting the experimental data to obtain impedance parameters. (<b>a</b>) without diffusion consideration. (<b>b</b>) with diffusion by adding Zw in the circuit. (<b>c</b>) to simulate corrosion product effect by adding inductor (L) in the circuit.</p>
Full article ">Figure 9
<p>Nyquist and Bode plots of Al, Al-10% Al<sub>2</sub>O<sub>3</sub>, Al-0.25% GO, and Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25 GO nanocomposites tested in (<b>a</b>,<b>e</b>) 0.1M NaCl, (<b>b</b>,<b>f</b>) 0.3M NaCl, (<b>c</b>,<b>g</b>) 0.6M NaCl, (<b>d</b>,<b>h</b>) 0.5M H<sub>2</sub>SO<sub>4</sub> electrolytes at room temperature.</p>
Full article ">Figure 10
<p>(<b>a</b>) Charge transfer resistance (R<sub>ct</sub>) and (<b>b</b>) pore resistance (R<sub>po</sub>) of Al, Al-10% Al<sub>2</sub>O<sub>3</sub>, Al-0.25% GO, and Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25 GO nanocomposites tested in different test solutions at room temperatures.</p>
Full article ">Figure 11
<p>Electrochemical noise analysis (ENA) of Al, Al−10% Al<sub>2</sub>O<sub>3</sub>, Al−0.25% GO, A−10% Al<sub>2</sub>O<sub>3</sub>−0.25% GO in 0.6 M NaCl. (<b>a</b>) potential time record, (<b>b</b>) current time record, Fast Fourier Transformation (FFT) PSD of- (<b>c</b>) Potential, (<b>d</b>) Current, (<b>e</b>) Resistance (V/I), (<b>f</b>) Resistance (V/I) for Al-0.25% GO.</p>
Full article ">Figure 12
<p>PDP plots of Al, Al−10% Al<sub>2</sub>O<sub>3</sub>, Al−0.25% GO, and Al−10% Al<sub>2</sub>O<sub>3</sub>−0.25 GO nanocomposites tested in (<b>a</b>) 0.1 M NaCl, (<b>b</b>) 0.3 M NaCl, (<b>c</b>) 0.6 M NaCl, (<b>d</b>) 0.5 M H<sub>2</sub>SO<sub>4</sub> electrolytes at room temperature.</p>
Full article ">Figure 13
<p>Corrosion rates comparison of Al, Al-10% Al<sub>2</sub>O<sub>3</sub>, Al-0.25% GO, and Al-10% Al<sub>2</sub>O<sub>3</sub>-0.25% GO nanocomposites samples in different test solutions.</p>
Full article ">Figure 14
<p>SEM images and EDS mapping of the post-corrosion surface morphology of Al-0.25% GO nanocomposite tested in 0.6 M NaCl.</p>
Full article ">Figure 15
<p>Schematic illustration of chlorides or sulfate ions diffusion mechanism penetrating (<b>a</b>) Al matrix, (<b>b</b>) graphene oxide (GO)-reinforced Al MMCs.</p>
Full article ">Figure 16
<p>SEM images and EDS mapping of the fractured Al-0.25% GO nanocomposite surface reveal the distribution of GO at the grain boundaries of the Al matrix.</p>
Full article ">Figure 17
<p>The 3D-optical surface profiling of post corrosion grooves developed on (<b>a</b>) Al, (<b>b</b>) Al-Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>) Al-GO, (<b>d</b>) Al-Al<sub>2</sub>O<sub>3</sub>-GO (hybrid-) nanocomposites tested in 0.6 M NaCl.</p>
Full article ">Figure 18
<p>Graphical illustration of average surface roughness and pitting depth obtained by optical profilometer of aluminum and its composites.</p>
Full article ">
15 pages, 3068 KiB  
Article
Wettability of a Polymethylmethacrylate Surface by Fluorocarbon Surfactant Solutions
by Fei Yan, Cheng Ma, Qingtao Gong, Zhiqiang Jin, Wangjing Ma, Zhicheng Xu, Lei Zhang and Lu Zhang
Chemistry 2024, 6(5), 1063-1077; https://doi.org/10.3390/chemistry6050061 (registering DOI) - 16 Sep 2024
Viewed by 94
Abstract
To clarify the adsorption behavior of fluorocarbon surfactants on PMMA surfaces, the contact angles of two nonionic fluorocarbon surfactants (FNS-1 and FNS-2) and an anionic fluorocarbon surfactant (FAS) on polymethylmethacrylate (PMMA) surface were determined using the sessile drop method. Moreover, the effects of [...] Read more.
To clarify the adsorption behavior of fluorocarbon surfactants on PMMA surfaces, the contact angles of two nonionic fluorocarbon surfactants (FNS-1 and FNS-2) and an anionic fluorocarbon surfactant (FAS) on polymethylmethacrylate (PMMA) surface were determined using the sessile drop method. Moreover, the effects of molecular structures on the surface tension, adhesion tension, solid–liquid interfacial tension, and adhesion work of the three fluorocarbon surfactants were investigated. The results demonstrate that the adsorption amounts for three fluorocarbon surfactants at the air–water interface are 4~5 times higher than those at the PMMA–solution interface. The three fluorocarbon surfactants adsorb on the PMMA surface by polar groups before CMC and by hydrophobic chains after CMC. Before CMC, FNS-2 with the smallest molecular size owns the highest adsorption amount, while FAS with large-branched chains and electrostatic repulsion has the smallest adsorption amount. After CMC, the three fluorocarbon surfactants form aggregates at the PMMA-liquid interface. FAS possesses the smallest adsorption amount after CMC. Besides, FNS-1 possesses a higher adsorption amount than FNS-2 due to the longer fluorocarbon chain and the lower CMC value of FNS-1. The adsorption behaviors of nonionic and anionic fluorocarbon surfactants on the PMMA surface are different. FAS forms interfacial aggregates before CMC, which may be attributed to the electrostatic interaction between the anionic head of FAS and the PMMA surface. Full article
(This article belongs to the Section Chemistry of Materials)
Show Figures

Figure 1

Figure 1
<p>Effect of fluorocarbon surfactant concentration on the surface tension.</p>
Full article ">Figure 2
<p>Effect of fluorocarbon surfactant concentration on the contact angles.</p>
Full article ">Figure 3
<p>The adhesion tension of three fluorocarbon surfactants varies with surface tension on the PMMA surface.</p>
Full article ">Figure 4
<p>Effect of fluorocarbon surfactant concentration on the PMMA–water interface tension.</p>
Full article ">Figure 5
<p>Effect of fluorocarbon surfactant concentration on adhesion work.</p>
Full article ">Figure 6
<p>Concentration dependence of adsorption parameters for fluorocarbon surfactants on PMMA surface.</p>
Full article ">Figure 7
<p>Adsorption mechanism of the fluorocarbon surfactants on PMMA surface.</p>
Full article ">Scheme 1
<p>Structures and abbreviations of the three fluorocarbon surfactants.</p>
Full article ">
22 pages, 7318 KiB  
Article
Multidimensional Characterization and Separation of Ultrafine Particles: Insights and Advances by Means of Froth Flotation
by Johanna Sygusch and Martin Rudolph
Powders 2024, 3(3), 460-481; https://doi.org/10.3390/powders3030025 (registering DOI) - 15 Sep 2024
Viewed by 240
Abstract
Particle systems and their efficient and precise separation are becoming increasingly complex. Therefore, instead of focusing on a single separation feature, a multidimensional approach is needed where more than one particle property is considered. This, however, requires the precise characterization of the particle [...] Read more.
Particle systems and their efficient and precise separation are becoming increasingly complex. Therefore, instead of focusing on a single separation feature, a multidimensional approach is needed where more than one particle property is considered. This, however, requires the precise characterization of the particle system, which is especially challenging for fine particles with sizes below 10 µm. This paper discusses the benefits and limitations of different characterization techniques, including optical contour analysis, inverse gas chromatography, flow cytometry, and SEM-based image analysis. The separation of ultrafine particles was investigated for a binary system using froth flotation, where a novel developed flotation apparatus is used. A special focus was placed on the multidimensional evaluation of the separation according to the particle properties of size, shape, and wettability, which was addressed via multivariate Tromp and entropy functions. The results emphasize the intricacy of the flotation process and the complex interaction of the individual particle properties and process parameters. The investigations contribute to the understanding of the characterization of particulate properties as well as the separation behavior of ultrafine particles via froth flotation, especially in the case of a multidimensional approach. Full article
Show Figures

Figure 1

Figure 1
<p>Scanning electron microscopy images of glass spheres (<b>left</b>), glass fragments (<b>middle</b>), and magnetite (<b>right</b>) used as feed.</p>
Full article ">Figure 2
<p>Schematic diagram of the <span class="html-italic">MultiDimFlot</span> separation apparatus (<b>left</b>) and the actual lab set-up (<b>right</b>) [<a href="#B38-powders-03-00025" class="html-bibr">38</a>].</p>
Full article ">Figure 3
<p>Cumulative frequency distributions of static, receding (rec), and advancing (adv) contact angles measured via optical contour analysis on glass slides with wettability states: hydrophilic C0 (red), moderately hydrophobic C6 (blue), and strongly hydrophobic C10 (green). The difference between the receding and the advancing contact angle represents the hysteresis.</p>
Full article ">Figure 4
<p>The energy of interaction of a particle with a hyperhydrophobic bubble in water ΔG<sub>pwb</sub> for magnetite (black hexagon) and glass spheres (blue circle) and fragments (green diamond) in different wettability states: pristine, unesterified C0; particles esterified with 1-hexanol C6; and particles esterified with 1-decanol C10. All values are calculated for surface coverages <span class="html-italic">n/nm</span> of 0.1% of probe molecules. The error bars are obtained via error propagation considering the involved regressions, and the lines are added to guide the eye.</p>
Full article ">Figure 5
<p>Example of a false-color image (subset) of glass spheres (red) and magnetite particles (blue) obtained using MLA with a resolution of 0.25 µm per pixel.</p>
Full article ">Figure 6
<p>Bivariate probability densities representing the particle descriptors of shape and size as aspect ratio and area-equivalent diameter, respectively, for glass spheres (<b>left</b>), glass fragments (<b>middle</b>), and magnetite (<b>right</b>). Their computation is based on the copula-based approach outlined in [<a href="#B37-powders-03-00025" class="html-bibr">37</a>] using MLA images of the individual fractions. The color scale indicates the frequency of the described property value. Particles with fewer than 4 pixels are excluded from the analysis of MLA images.</p>
Full article ">Figure 7
<p>Results of glass spheres (<b>left</b>), glass fragments (<b>middle</b>), and magnetite (<b>right</b>) obtained via flow cytometry, with sideward scattering (SSC) and forward scattering (FSC) holding information on particle shape and size, respectively. The scattering values represent the different detection channels and have no units. The color intensity indicates the frequency of the described property value.</p>
Full article ">Figure 8
<p>Fuerstenau upgrading curves for the kinetic flotation tests of glass spheres (<b>left</b>) and fragments (<b>right</b>) of three different wettability states: C0 hydrophilic (red); C6 moderately hydrophobic (blue); C10 strongly hydrophobic (green), where every fraction is mixed with magnetite as the hydrophilic feed. Single data points represent individual test runs (T), whereas filled data points, connected by a dashed line to guide the eye, represent average values. Each data point corresponds to the cumulative recovery after defined flotation times.</p>
Full article ">Figure 9
<p>Fuerstenau upgrading curves for the kinetic flotation tests of glass spheres (black circles) and fragments (pink diamonds) of three different wettability states: C0 hydrophilic (<b>left</b>); C6 moderately hydrophobic (<b>middle</b>); C10 strongly hydrophobic (<b>right</b>), where every fraction is mixed with magnetite as the hydrophilic feed. Single data points represent individual test runs (T), whereas filled data points, connected by a dashed line to guide the eye, represent average values. Each data point corresponds to the cumulative recovery after defined flotation times.</p>
Full article ">Figure 10
<p>Bivariate entropy functions for glass spheres (<b>upper row</b>) and glass fragments (<b>lower row</b>) with different wettability states: unesterified hydrophilic C0 (<b>left</b>), moderately hydrophobic C6 (<b>middle</b>), and strongly hydrophobic C10 (<b>right</b>). All glass particle fractions are mixed with magnetite as feed material. The color code indicates the value for the entropy function <span class="html-italic">H</span>.</p>
Full article ">Figure 11
<p>Bivariate entropy functions for magnetite exclusively, mixed with glass spheres (<b>upper row</b>) and glass fragments (<b>lower row</b>) with different wettability states: unesterified hydrophilic C0 (<b>left</b>), moderately hydrophobic C6 (<b>middle</b>), and strongly hydrophobic C10 (<b>right</b>). The color code indicates the value for the entropy function <span class="html-italic">H</span>, where dark green represents a separation uncertainty of zero.</p>
Full article ">
14 pages, 4069 KiB  
Article
Electroless ZnO Deposition on Mg-Al Alloy for Improved Corrosion Resistance to Marine Environments
by Luis Chávez, Lucien Veleva and Andrea Castillo-Atoche
Coatings 2024, 14(9), 1192; https://doi.org/10.3390/coatings14091192 - 15 Sep 2024
Viewed by 190
Abstract
Electroless ZnO (≈900 nm) was deposited on the surface of an Mg-Al alloy (AM60) to reduce its degradation in the marine environment. Uncoated and coated ZnO samples were exposed to an SME simulated marine solution for up to 30 days. The AFM and [...] Read more.
Electroless ZnO (≈900 nm) was deposited on the surface of an Mg-Al alloy (AM60) to reduce its degradation in the marine environment. Uncoated and coated ZnO samples were exposed to an SME simulated marine solution for up to 30 days. The AFM and optical images revealed that the corrosion attack on the ZnO-AM60 surface was reduced due to an increase in the surface hydrophobicity of the ZnO coating (contact angle of ≈91.6°). The change in pH to more alkaline values over time was less pronounced for ZnO-AM60 (by ≈13%), whereas the release of Mg2+ ions was reduced by 34 times, attributed to the decrease in active sites on the Mg-matrix provided by the electroless ZnO coating. The OCP (free corrosion potential) of ZnO-AM60 shifted towards less negative values of ≈100 mV, indicating that electroless ZnO may serve as a good barrier for AM60 in a marine environment. The calculated polarization resistance (Rp), based on EIS data, was ≈3 times greater for ZnO-AM60 than that of the uncoated substrate. Full article
(This article belongs to the Special Issue Surface Modification of Magnesium, Aluminum Alloys, and Steel)
Show Figures

Figure 1

Figure 1
<p>2D AFM micrographs of (<b>a</b>) AM60 and (<b>b</b>) ZnO-AM60 surfaces.</p>
Full article ">Figure 2
<p>Thickness of the ZnO electroless coating deposited on AM60 alloy surface.</p>
Full article ">Figure 3
<p>Optical images of (<b>a</b>) sample pattern (ZnO-AM60) and after adhesion test with a force of (<b>b</b>) 0.1 mN, (<b>c</b>) 0.05 mN and (<b>d</b>) 0.01 mN.</p>
Full article ">Figure 4
<p>XPS depth profile analysis of ZnO-AM60.</p>
Full article ">Figure 5
<p>High resolution XPS spectra of: (<b>a</b>) Zn2p, (<b>b</b>) O1s, (<b>c</b>) Mg1s, and (<b>d</b>) C1s obtained after 60s of sputtering (erosion) time of the ZnO electroless coating deposited on AM60 alloy.</p>
Full article ">Figure 6
<p>Measured contact angle values of uncoated (<b>a</b>) AM60 and (<b>b</b>) ZnO-AM60.</p>
Full article ">Figure 7
<p>Optical images (×50) of: (<b>a</b>) uncoated AM60 and (<b>b</b>) ZnO-AM60 after exposure for 30 days to the aggressive SME solution.</p>
Full article ">Figure 8
<p>(<b>a</b>) EIS Nyquist diagrams and Bode impedance (<b>b</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mfenced open="|" close="|" separators="|"> <mrow> <mi mathvariant="normal">Z</mi> </mrow> </mfenced> </mrow> <mrow> <mrow> <mn>0.01</mn> <mi>Hz</mi> </mrow> </mrow> </msub> </mrow> </semantics></math> at low frequency of uncoated AM60 and ZnO-AM60 as a function of the immersion time in a simulated marine environment (SME).</p>
Full article ">Figure 9
<p>Equivalent circuit for fitting EIS experimental data of AM60 and ZnO-AM60 during their exposure to SME solution.</p>
Full article ">
19 pages, 7344 KiB  
Review
Patterning of Organic Semiconductors Leads to Functional Integration: From Unit Device to Integrated Electronics
by Wangmyung Choi, Yeo Eun Kim and Hocheon Yoo
Polymers 2024, 16(18), 2613; https://doi.org/10.3390/polym16182613 - 15 Sep 2024
Viewed by 180
Abstract
The use of organic semiconductors in electronic devices, including transistors, sensors, and memories, unlocks innovative possibilities such as streamlined fabrication processes, enhanced mechanical flexibility, and potential new applications. Nevertheless, the increasing technical demand for patterning organic semiconductors requires greater integration and functional implementation. [...] Read more.
The use of organic semiconductors in electronic devices, including transistors, sensors, and memories, unlocks innovative possibilities such as streamlined fabrication processes, enhanced mechanical flexibility, and potential new applications. Nevertheless, the increasing technical demand for patterning organic semiconductors requires greater integration and functional implementation. This paper overviews recent efforts to pattern organic semiconductors compatible with electronic devices. The review categorizes the contributions of organic semiconductor patterning approaches, such as surface-grafting polymers, capillary force lithography, wettability, evaporation, and diffusion in organic semiconductor-based transistors and sensors, offering a timely perspective on unconventional approaches to enable the patterning of organic semiconductors with a strong focus on the advantages of organic semiconductor utilization. In addition, this review explores the opportunities and challenges of organic semiconductor-based integration, emphasizing the issues related to patterning and interconnection. Full article
(This article belongs to the Special Issue Polymer-Based Smart Materials: Preparation and Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of visible light-mediated p(DMAEMA) brush growth under ambient conditions. (<b>b</b>) Optical image of photomask-guided sequential p(DMAEMA) growth on a wafer-based substrate patterned with underlying polymers (X: substrate, Y: patterned polymer). Atomic force microscopy topography images and height profiles for patterned p(DMAEMA), at (<b>c</b>) X region (substrate) and (<b>d</b>) Y region (patterned polymer). The p(DMAEMA) brush synthesis process under ambient conditions using M, R, and L letter glass covers: (<b>e</b>) initial chemically treated silicon wafer, (<b>f</b>) polymer brush growth process through M, R, and L letter covers, (<b>g</b>) p(DMAEMA) brushes patterned with the letters M, R, and L. (<b>h</b>) Enlarged optical image of p(DMAEMA) brushes patterned in the shape of MRL [<a href="#B54-polymers-16-02613" class="html-bibr">54</a>]. Copyright © 2018, John Wiley and Sons. (<b>i</b>) Schematic diagram of the PS-brush synthesis process. (<b>j</b>) Schematic diagram of the evaporation process for organic semiconductor crystal growth between the PS brushes and organic semiconductor ink. (<b>k</b>) Mechanism of organic semiconductor crystal growth depending on the chain length of PS brushes. (<b>l</b>) Schematic diagram of the device and transfer curve of the TIPS-pentacene-based TFT with optimized crystallinity [<a href="#B57-polymers-16-02613" class="html-bibr">57</a>]. Copyright © 2016, John Wiley and Sons.</p>
Full article ">Figure 2
<p>Capillary force-based (<b>a</b>) schematic diagram of the C8-BTBT patterning process, (<b>b</b>) sequential optical microscopy images of the formation process of C8-BTBT crystals (i: solid; ii and iii: phase transition; iv: microchannel formation; v: cooling), and (<b>c</b>) COMSOL multiphysics simulation of liquid C8-BTBT molecule flow (i: C8-BTBT deposition; ii: C8-BTBT liquid moves along the sidewall of CYTOP; iii: C8-BTBT liquid starts to fill the bottoms; iv and v: as the C8-BTBT liquid fills the droplets form an upward concave meniscus; vi and vii: C8-BTBT liquid fills the voids due to capillary forces; viii: C8-BTBT microchannel formation). (<b>d</b>) Schematic diagram of a 13 × 13 array of OTFTs based on single-crystal C8-BTBT. At the unit device scale, (<b>e</b>,<b>f</b>) scanning electron microscopy images of a single-crystal C8-BTBT-based OFET and (<b>g</b>) transfer curve (red line: logarithmic scale; purple line: linear scale; dashed line: leakage current) [<a href="#B55-polymers-16-02613" class="html-bibr">55</a>]. Copyright © 2020, Elsevier. 1D single-crystal C8-BTBT array synthesis: (<b>h</b>) schematic diagram and (<b>i</b>) mechanism. (<b>j</b>) Transfer curves of OFETs based on single-crystal C8-BTBT morphology (left OFET: 1D single-crystalline belt arrays; right OFET: thin film; red line: logarithmic scale; blue line: linear scale). 1D single-crystal C8-BTBT-based pressure sensor: (<b>k</b>) schematic diagram, (<b>l</b>) photograph of the device, and (<b>m</b>) pressure–current response curves [<a href="#B60-polymers-16-02613" class="html-bibr">60</a>]. Copyright © 2018, John Wiley and Sons.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic diagram showing the patterning process for C8-BTBT LC films using the IJP-MP method. (<b>b</b>) Polarized optical microscopy image of a 7 × 7 OFET array patterned with C8-BTBT LC film. (<b>c</b>) Polarized optical microscopy images of a typical OFET device. (<b>d</b>) Transfer characteristics measured from 49 OFETs within the 7 × 7 device array. (<b>e</b>) Spatial distribution of saturation mobilities across the 49 OFET devices [<a href="#B51-polymers-16-02613" class="html-bibr">51</a>]. Copyright © 2021, John Wiley and Sons. (<b>f</b>) Schematic diagram showing the fabrication of patterned OSC crystals on a wettability-controlled substrate using the bar-coating method. The inset details the pattern design details and an enlarged view of the film crystallization processes. (<b>g</b>) Cross-sectional and top views of the meniscus and contact line at the interface between the wetting (white) and dewetting (light blue) regions with high (Δ<span class="html-italic">γ</span> ≥ 31.5 mN·m<sup>−1</sup>) and appropriate (Δ<span class="html-italic">γ</span> = 21.1 mN·m<sup>−1</sup>) surface tension differences [<a href="#B58-polymers-16-02613" class="html-bibr">58</a>]. Copyright © 2021, John Wiley and Sons. (<b>h</b>) Polarized optical microscopy images of coated films on substrates with large and small surface tension differences. Preparation of a superhydrophobic surface on a silicon wafer using atmospheric RF plasma, followed by the formation of a selective hydrophilic area using an EHD jet printing system. (<b>i</b>) fluorescent image of polystyrene nanoparticles after deposition (scale bar: 50 μm) [<a href="#B106-polymers-16-02613" class="html-bibr">106</a>]. Copyright © 2017, Elsevier.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic diagram of the formation process of honeycomb-patterned films. (<b>b</b>) Large-area honeycomb film for digital photography. (<b>c</b>) Scanning electron microscopy image of the honeycomb-patterned film. (<b>d</b>) Scanning electron microscopy images of the surfaces of patterned films made from different polymers, along with the molecular structural formulas of these polymers: PS, PSF, and PES [<a href="#B116-polymers-16-02613" class="html-bibr">116</a>]. Copyright © 2021, American Chemical Society. (<b>e</b>) Illustration of the sphere-on-flat evaporation setup. (<b>f</b>) Illustration of the concentric rings of as-prepared organic nanowires. (<b>g</b>) Optical micrographs and (<b>h</b>) scanning electron microscopy images of concentric rings of DMQA nanowires formed during solvent evaporation [<a href="#B117-polymers-16-02613" class="html-bibr">117</a>]. Copyright © 2011, John Wiley and Sons.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic diagram of the preparation procedure for an array of silica pillars. (<b>b</b>) Representative scanning electron microscopy image of the patterned silica pillar array on the surface [<a href="#B137-polymers-16-02613" class="html-bibr">137</a>]. Copyright © 2015, American Chemical Society. (<b>c</b>) Chemical structures of pNaSS. Illustration of the molecular gate concept through PL switching in PFO films, depicted for (<b>d</b>) thermal and (<b>e</b>) solvent vapor-based stimuli. The top rows show photographs of the films under UV light at various processing steps, with schematic illustrations in the bottom rows. (<b>f</b>) Chemical structures of PFO (R = C<sub>8</sub>H<sub>17</sub>). (<b>g</b>) Schematic diagram of the laser-patterning process. (<b>h</b>) PL intensity at 438 nm showing a spectral selection of β-phase emission. In addition, a Raman intensity ratio image (right panel) of the line patterned using <span class="html-italic">P</span> = 30 mW was obtained, indicating the local β-phase fraction as estimated from the ratio of Raman intensities at 1257 and 1606 cm<sup>−1</sup> [<a href="#B138-polymers-16-02613" class="html-bibr">138</a>]. Copyright © 2020, Springer Nature.</p>
Full article ">
16 pages, 12097 KiB  
Article
Insight into the Structural and Performance Correlation of Photocatalytic TiO2/Cu Composite Films Prepared by Magnetron Sputtering Method
by Kun Lu, Miao Sun, Yaohong Jiang, Xinmeng Wu, Lijun Zhao and Junhua Xu
Catalysts 2024, 14(9), 621; https://doi.org/10.3390/catal14090621 - 14 Sep 2024
Viewed by 314
Abstract
Photocatalysis technology, as an efficient and safe environmentally friendly purification technique, has garnered significant attention and interest. Traditional TiO2 photocatalytic materials still face limitations in practical applications, hindering their widespread adoption. The research prepared TiO2/Cu films with different Cu contents [...] Read more.
Photocatalysis technology, as an efficient and safe environmentally friendly purification technique, has garnered significant attention and interest. Traditional TiO2 photocatalytic materials still face limitations in practical applications, hindering their widespread adoption. The research prepared TiO2/Cu films with different Cu contents using a magnetron sputtering multi-target co-deposition technique. The incorporation of Cu significantly enhances the antibacterial properties and visible light response of the films. The effects of different Cu contents on the microstructure, surface morphology, wettability, antibacterial properties, and visible light response of the films were investigated using an X-ray diffractometer, X-ray photoelectron spectrometer, field emission scanning electron microscope, confocal laser scanning microscope, Ultraviolet–visible spectrophotometer, and contact angle goniometer. The results showed that the prepared TiO2/Cu films were mainly composed of the rutile TiO2 phase and face-center cubic Cu phase. The introduction of Cu affected the crystal orientation of TiO2 and refined the grain size of the films. With the increase in Cu content, the surface roughness of the films first decreased and then increased. The water contact angle of the films first increased and then decreased, and the film exhibited optimal hydrophobicity when the Cu target power was 10 W. The TiO2/Cu films showed good antibacterial properties against Escherichia coli and Staphylococcus aureus. The introduction of Cu shifted the absorption edge of the films to the red region, significantly narrowed the band gap width to 2.5 eV, and broadened the light response range of the films to the visible light region. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of TiO<sub>2</sub>/Cu composite films with different Cu target powers.</p>
Full article ">Figure 2
<p>XPS spectra of TiO<sub>2</sub>/Cu composite films with Cu target power of 15 W (<b>a</b>) full spectra. (<b>b</b>) Ti 2p. (<b>c</b>) Cu 2p and (<b>d</b>) O 1s.</p>
Full article ">Figure 3
<p>Surface morphology of TiO<sub>2</sub>/Cu composite films after annealing at 500 °C (<b>a</b>) TiO<sub>2</sub>; (<b>b</b>) Cu10; (<b>c</b>) Cu15; (<b>d</b>) Cu20.</p>
Full article ">Figure 4
<p>Laser confocal spectra of TiO<sub>2</sub>/Cu composite films with different Cu target powers (<b>a</b>) TiO<sub>2</sub>; (<b>b</b>) Cu10; (<b>c</b>) Cu15; (<b>d</b>) Cu20.</p>
Full article ">Figure 5
<p>Contact angle of TiO<sub>2</sub>/Cu films with different Cu target power (<b>a</b>) TiO<sub>2</sub>; (<b>b</b>) Cu10; (<b>c</b>) Cu15; (<b>d</b>) Cu20.</p>
Full article ">Figure 6
<p>Contact angle of TiO<sub>2</sub>/Cu films with different Cu target power.</p>
Full article ">Figure 7
<p>Antibacterial effect of TiO<sub>2</sub>/Cu composite film (<b>a</b>,<b>b</b>) <span class="html-italic">Escherichia coli</span>, (<b>c</b>,<b>d</b>) <span class="html-italic">Staphylococcus aureus</span>.</p>
Full article ">Figure 8
<p>(<b>a</b>) UV–Vis diffuse reflectance spectra and (<b>b</b>) optical band gap of TiO<sub>2</sub>/Cu composite films with different Cu target power.</p>
Full article ">Figure 9
<p>The antibacterial mechanism of TiO<sub>2</sub>/Cu composite film.</p>
Full article ">Figure 10
<p>Three-dimensional model of magnetron sputtering system.</p>
Full article ">Figure 11
<p>The schematic diagram of antibacterial test by coating plate method.</p>
Full article ">
21 pages, 3419 KiB  
Article
Novel Bioplastic Based on PVA Functionalized with Anthocyanins: Synthesis, Biochemical Properties and Food Applications
by Giuseppe Tancredi Patanè, Antonella Calderaro, Stefano Putaggio, Giovanna Ginestra, Giuseppina Mandalari, Santa Cirmi, Davide Barreca, Annamaria Russo, Teresa Gervasi, Giovanni Neri, Meryam Chelly, Annamaria Visco, Cristina Scolaro, Francesca Mancuso, Silvana Ficarra, Ester Tellone and Giuseppina Laganà
Int. J. Mol. Sci. 2024, 25(18), 9929; https://doi.org/10.3390/ijms25189929 (registering DOI) - 14 Sep 2024
Viewed by 268
Abstract
Over the last ten years, researchers’ efforts have aimed to replace the classic linear economy model with the circular economy model, favoring green chemical and industrial processes. From this point of view, biologically active molecules, coming from plants, flowers and biomass, are gaining [...] Read more.
Over the last ten years, researchers’ efforts have aimed to replace the classic linear economy model with the circular economy model, favoring green chemical and industrial processes. From this point of view, biologically active molecules, coming from plants, flowers and biomass, are gaining considerable value. In this study, firstly we focus on the development of a green protocol to obtain the purification of anthocyanins from the flower of Callistemon citrinus, based on simulation and on response surface optimization methodology. After that, we utilize them to manufacture and add new properties to bioplastics belonging to class 3, based on modified polyvinyl alcohol (PVA) with increasing amounts from 0.10 to 1.00%. The new polymers are analyzed to monitor morphological changes, optical properties, mechanical properties and antioxidant and antimicrobial activities. Fourier transform infrared spectroscopy (FTIR) spectra of the new materials show the characteristic bands of the PVA alone and a modification of the band at around 1138 cm−1 and 1083 cm−1, showing an influence of the anthocyanins’ addition on the sequence with crystalline and amorphous structures of the starting materials, as also shown by the results of the mechanical tests. These last showed an increase in thickening (from 29.92 μm to approx. 37 μm) and hydrophobicity with the concomitant increase in the added anthocyanins (change in wettability with water from 14° to 31°), decreasing the poor water/moisture resistance of PVA that decreases its strength and limits its application in food packaging, which makes the new materials ideal candidates for biodegradable packaging to extend the shelf-life of food. The functionalization also determines an increase in the opacity, from 2.46 to 3.42 T%/mm, the acquisition of antioxidant activity against 2,2-diphenyl-1-picrylhdrazyl and 2,2′-azino-bis-(3-ethylbenzothiazoline-6-sulfonic acid) radicals and, in the ferric reducing power assay, the antimicrobial (bactericidal) activity against different Staphylococcus aureus strains at the maximum tested concentration (1.00% of anthocyanins). On the whole, functionalization with anthocyanins results in the acquisition of new properties, making it suitable for food packaging purposes, as highlighted by a food fresh-keeping test. Full article
(This article belongs to the Section Materials Science)
Show Figures

Figure 1

Figure 1
<p>Representative pictogram for the green index of the extraction of anthocyanin from <span class="html-italic">Callistemon citrinus</span> using two different methods. (<b>A</b>) The index for the extraction with accelerated solvent extraction using methanol solution; (<b>B</b>) The index for the extraction with MAE using the new ethanol (EtOH) solution. In both pictograms, the colour scale (red-yellow-green) indicates the performance at each stage of the procedure. The less chemically ‘green’ the process, the more red it appears.</p>
Full article ">Figure 2
<p>Pareto chart diagram for the extraction of total anthocyanin content (TAC) from <span class="html-italic">Callistemon citrinus</span>. A = microwave power (W); B = extraction time (min); C = EtOH% in the extraction solution.</p>
Full article ">Figure 3
<p>Response surface plots analysis for the total anthocyanin content yield (TAC) from <span class="html-italic">Callistemon citrinus</span> powder with microwave assisted extraction (MAE) with the same solid-liquid ratio, 1:10 (<span class="html-italic">w</span>/<span class="html-italic">v</span>). (<b>A</b>) microwave power and EtOH % in the reaction mix; (<b>B</b>) microwave power and minutes; (<b>C</b>) EtOH % in the reaction mix and minutes.</p>
Full article ">Figure 4
<p>Representative FTIR spectra of the new PVA-based bioplastics produced by the addition of increasing amount of anthocyanins (0.0–1.0%). (<b><span style="color:lime">―</span></b>) Anthocyanins powder; (<b>―</b>) PVA alone; (<b><span style="color:red">―</span></b>) PVA plus 0.1% anthocyanins; (<b><span style="color:#2E74B5">―</span></b>) PVA plus 0.25% anthocyanins; (<b><span style="color:#FF66FF">―</span></b>) PVA plus 0.5% anthocyanins; (<b><span style="color:#66FFFF">―</span></b>) PVA plus 1.0% anthocyanins.</p>
Full article ">Figure 5
<p>Release for short- and long-term migration of anthocyanins from PVA films in different food simulants. (<b>A</b>) PVA plus 0.10% of anthocyanins; (<b>B</b>) PVA plus 0.25% of anthocyanins; (<b>C</b>) PVA plus 0.50% of anthocyanins; (<b>D</b>) PVA plus 1.00% of anthocyanins. (<span style="color:#3366CC">⬤</span>) H<sub>2</sub>O; (<span style="color:red">■</span>) ethanol 10%; (<span style="color:green">▲</span>) ethanol 50%; (<span style="color:#7100E2">▼</span>) acetic acid 3%.</p>
Full article ">Figure 6
<p>Evaluation of antioxidant activity of PVA-based bioplastics with different % of anthocyanins (0.10, 0.25, 0.50, 1%) in the most common antioxidant assays. (<b>A</b>) ABTS assay; (<b>B</b>) DPPH assay; (<b>C</b>) ferric reducing power (FRAP) assay. The letters in the different graph indicate: a, control sample; b, PVA alone; c, PVA plus 0.10% of anthocyanins; d, PVA plus 0.25% of anthocyanins; e, PVA plus 0.50% of anthocyanins; f, PVA plus 1.00% of anthocyanins. *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Preparation of bags for food packaging produced with PVA plus 1.0% of anthocyanins and its utilization for apple samples.</p>
Full article ">Figure 8
<p>Analysis of the changes in apple samples packed or not packed after different intervals of time (0 and 72 h). (<b>A</b>) Changes in the browning of the samples monitored at 420 nm. The letters in the graph indicate: a, apple samples not packed after 0 h; b, apple samples not packed after 72 h; c, apple sample packed with PVA film alone after 72 h; d, apple samples packed with PVA plus 1.00% of anthocyanins after 72 h. (<b>B</b>) Changes in the antioxidant potential monitored by DPPH assay. The letters in the graph indicate: a, control without samples; b, apple samples not packed after 0 h; c, apple samples not packed after 72 h; d, apple sample packed with PVA film alone after 72 h; e, apple samples packed with PVA plus 1.00% of anthocyanins after 72 h. The ** indicates significant changes with respect to the control at <span class="html-italic">p</span> &gt; 0.05.</p>
Full article ">
23 pages, 17525 KiB  
Article
The Preparation and Properties of a Ni-SiO2 Superamphiphobic Coating Obtained by Electrodeposition
by Jianguo Liu, Songlin Tong, Shuaihua Wang, Zhiyao Wan, Xiao Xing and Gan Cui
Metals 2024, 14(9), 1047; https://doi.org/10.3390/met14091047 - 14 Sep 2024
Viewed by 194
Abstract
Superamphiphobic coatings have shown great potential in many fields such as with their anti-corrosion, high-temperature resistance, self-cleaning, and drag reduction properties. However, due to the poor stability of their coatings, it is difficult to apply them on a large scale. In this paper, [...] Read more.
Superamphiphobic coatings have shown great potential in many fields such as with their anti-corrosion, high-temperature resistance, self-cleaning, and drag reduction properties. However, due to the poor stability of their coatings, it is difficult to apply them on a large scale. In this paper, two kinds of SiO2 particles and nickel were co-deposited on the surface of steel to construct a micro/nano dual-scale structure by composite electrodeposition. The surface of the coating was then fluorinated with the low-surface-energy material 1H,1H,2H,2H-Perfluorodecyltriethoxysilane (AC-FAS) to prepare a Ni-SiO2 superamphiphobic coating. The coating has a water contact angle of 159° and an oil contact angle of 151°. The effect of nanoparticle concentration on the wettability and surface morphology of the coating was systematically studied. Comparative experiments revealed that the optimal micro/nanoparticle concentrations were 8 g/L of 20 nm SiO2 and 2 g/L of 1 μm SiO2. This preparation method greatly improves the corrosion resistance, wear resistance, chemical stability, and high-temperature resistance of the coating. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of electrodeposition double-electrode system.</p>
Full article ">Figure 2
<p>Schematic illustration of the abrasion test.</p>
Full article ">Figure 3
<p>The WCA and WSA of coatings at different concentrations of micro/nanoparticles. The error bars indicate standard deviations.</p>
Full article ">Figure 4
<p>The OCA and OSA of coatings at different concentrations of micro/nanoparticles. The error bars indicate standard deviations.</p>
Full article ">Figure 5
<p>SEM morphology of composite coatings under different particle size–particle concentration ratios: (<b>a</b>) 8/0, (<b>b</b>) 6/2, (<b>c</b>) 4/4, (<b>d</b>) 2/6, (<b>e</b>) 0/8.</p>
Full article ">Figure 6
<p>The SEM morphology of composite coatings at a concentration ratio of 6/2 of nanoscale particles to micron-sized particles.</p>
Full article ">Figure 7
<p>EDS diagram of superamphiphobic coatings.</p>
Full article ">Figure 8
<p>The distribution of elements on the surface of superamphiphobic coatings.</p>
Full article ">Figure 9
<p>The three-dimensional topography of the superamphiphobic coating’s surface.</p>
Full article ">Figure 10
<p>Cross-sectional morphology of superamphiphobic coatings.</p>
Full article ">Figure 11
<p>The change in the contact angle of the Ni-SiO<sub>2</sub> superamphiphobic coating with immersion time: (<b>a</b>) water contact angle, (<b>b</b>) oil contact angle.</p>
Full article ">Figure 12
<p>Polarization curves of Steel, Ni coating, and Ni-SiO<sub>2</sub> superamphiphobic coating.</p>
Full article ">Figure 13
<p>Nyquist plots of bare substrate, Ni coating, and Ni-SiO<sub>2</sub> superamphiphobic coating.</p>
Full article ">Figure 14
<p>Nyquist plots of steel immersed for different times.</p>
Full article ">Figure 15
<p>Bode plots of steel immersed for different times. (<b>a</b>) Frequency–impedance plot; (<b>b</b>) frequency–phase angle plot.</p>
Full article ">Figure 16
<p>Equivalent circuit of steel in solution.</p>
Full article ">Figure 17
<p>Nyquist plots of pure Ni coating immersed for different times.</p>
Full article ">Figure 18
<p>Bode plots of pure Ni coating immersed for different times. (<b>a</b>) Frequency–impedance plot; (<b>b</b>) frequency–phase angle plot.</p>
Full article ">Figure 19
<p>Equivalent circuits of Ni-SiO<sub>2</sub> superamphiphobic coating in 3.5 wt.% NaCl solution. (<b>a</b>) initial time, (<b>b</b>) failure time.</p>
Full article ">Figure 20
<p>Nyquist plots of Ni-SiO<sub>2</sub> superamphiphobic coating immersed for different times.</p>
Full article ">Figure 21
<p>Bode plots of Ni-SiO<sub>2</sub> superamphiphobic coating immersed for different times. (<b>a</b>) Frequency–impedance plot; (<b>b</b>) frequency–phase angle plot.</p>
Full article ">Figure 22
<p>A comparison of the hardness of steel, Ni coating, and Ni-SiO<sub>2</sub> superamphiphobic coating.</p>
Full article ">Figure 23
<p>Contact angle of pure Ni superhydrophobic coating under different wear distance: (<b>a</b>) water contact angle; (<b>b</b>) oil contact angle.</p>
Full article ">Figure 24
<p>Contact angle of Ni-SiO<sub>2</sub> composite superamphiphobic coating under different wear distance: (<b>a</b>) water contact angle; (<b>b</b>) oil contact angle.</p>
Full article ">Figure 25
<p>Surface morphology of samples after wear: (<b>a</b>) pure Ni coating; (<b>b</b>) Ni-SiO<sub>2</sub> composite superamphiphobic coating.</p>
Full article ">Figure 26
<p>The contact angle of the Ni-SiO<sub>2</sub> superamphiphobic coating at different temperatures.</p>
Full article ">Figure 27
<p>The SEM morphology of the Ni-SiO<sub>2</sub> superamphiphobic coating after high-temperature treatment at 300 °C.</p>
Full article ">Figure 28
<p>EDS diagram of Ni-SiO<sub>2</sub> superamphiphobic coating after high temperature treatment at 300 °C.</p>
Full article ">
14 pages, 5812 KiB  
Article
Partially Bio-Based and Biodegradable Poly(Propylene Terephthalate-Co-Adipate) Copolymers: Synthesis, Thermal Properties, and Enzymatic Degradation Behavior
by Ping Song, Mingjun Li, Haonan Wang, Yi Cheng and Zhiyong Wei
Polymers 2024, 16(18), 2588; https://doi.org/10.3390/polym16182588 - 13 Sep 2024
Viewed by 223
Abstract
A series of partially bio-based and biodegradable poly(propylene terephthalate-co-adipate) (PPTA) random copolymers with different components were prepared by the melt polycondensation of petro-based adipic acid and terephthalic acid with bio-based 1,3-propanediol. The microstructure, crystallization behavior, thermal properties, and enzymatic degradation properties were further [...] Read more.
A series of partially bio-based and biodegradable poly(propylene terephthalate-co-adipate) (PPTA) random copolymers with different components were prepared by the melt polycondensation of petro-based adipic acid and terephthalic acid with bio-based 1,3-propanediol. The microstructure, crystallization behavior, thermal properties, and enzymatic degradation properties were further investigated. The thermal decomposition kinetics was deeply analyzed using Friedman’s method, with the thermal degradation activation energy ranging from 297.8 to 302.1 kJ/mol. The crystallinity and wettability of the copolymers decreased with the increase in the content of the third unit, but they were lower than those of the homopolymer. The thermal degradation activation energy E, carbon residue, and reaction level n all showed a decreasing trend. Meanwhile, the initial thermal decomposition temperature (Td) was higher than 350 °C, which can meet the requirements for processing and use. The PPTA copolymer material still showed excellent thermal stability. Adding PA units could regulate the crystallinity, wettability, and degradation rate of PPTA copolymers. The composition of PPTA copolymers in different degradation cycles was characterized by 1H NMR analysis. Further, the copolymers’ surface morphology during the process of enzymatic degradation also was observed by scanning electron microscopy (SEM). The copolymers’ enzymatic degradation accorded with the surface degradation mechanism. The copolymers showed significant degradation behavior within 30 days, and the rate increased with increasing PA content when the PA content exceeded 45.36%. Full article
(This article belongs to the Special Issue Synthesis and Application of Degradable Polymers)
Show Figures

Figure 1

Figure 1
<p>Synthetic routine of PPTA copolymers.</p>
Full article ">Figure 2
<p>(<b>a</b>) <sup>1</sup>H NMR and (<b>b</b>) GPC spectra of PPTA copolymers.</p>
Full article ">Figure 3
<p>(<b>a</b>) Cooling curves and (<b>b</b>) the second heating curves for PPTA copolymers.</p>
Full article ">Figure 4
<p>The relative crystallinity of (<b>a</b>) PPTA-20 and (<b>b</b>) PPTA-40 changes with time at different isothermal crystallization temperatures. Avrami analysis by plotting <span class="html-italic">ln</span>(−<span class="html-italic">ln</span>(l − <span class="html-italic">X</span><sub>t</sub>)) vs. ln<span class="html-italic">t</span> of (<b>c</b>) PPTA-20 and (<b>d</b>) PPTA-40 at various <span class="html-italic">T</span><sub>c</sub> values.</p>
Full article ">Figure 5
<p>(<b>a</b>) Thermogravimetric analysis (TGA), (<b>b</b>) derivative thermogravimetry (DTG), (<b>c</b>) ln(dα/dt) vs. 10<sup>4</sup>/T curves, and (<b>d</b>) <span class="html-italic">ln</span>(1 − α) vs. 10<sup>4</sup>/T curves of PPTA with different compositions.</p>
Full article ">Figure 6
<p>Contact angle photographs of (<b>a</b>) PPT, (<b>b</b>) PPTA-20, (<b>c</b>) PPTA-40, (<b>d</b>) PPTA-60, (<b>e</b>) PPTA-80, and (<b>f</b>) PPA. (<b>g</b>) Diagram of contact angle changing with PA content. (<b>h</b>) Curves of weightlessness of PPTA with different compositions over time.</p>
Full article ">Figure 7
<p>SEM diagrams of PPTA surface under different times of enzyme degradation.</p>
Full article ">
5 pages, 4202 KiB  
Proceeding Paper
Computational Fluid Dynamics Analysis of an Innovative Multi-Purpose Green Roof
by Seyed Navid Naghib, Behrouz Pirouz, Hana Javadi Nejad, Michele Turco, Stefania Anna Palermo and Patrizia Piro
Eng. Proc. 2024, 69(1), 133; https://doi.org/10.3390/engproc2024069133 - 13 Sep 2024
Viewed by 78
Abstract
In this study, to improve the application and performance of conventional green roof systems, a novel multi-purpose green roof system was simulated numerically using computational fluid dynamics (CFD). The innovative multi-purpose green roof contains a soil layer and water filter, meaning the water [...] Read more.
In this study, to improve the application and performance of conventional green roof systems, a novel multi-purpose green roof system was simulated numerically using computational fluid dynamics (CFD). The innovative multi-purpose green roof contains a soil layer and water filter, meaning the water retention time not only depends on the soil media but also depends on the filter’s pore size, improving the impact on runoff quality and quantity. In this regard, after mesh sensitivity analysis, the developed model was validated using experimental data, and the results show the accuracy of CFD in the simulation of porous media and filters. Comparisons between experimental and numerical results demonstrate the impact of proper porosity values in the simulation of a porous environment and reveal the source of errors in the numerical prediction of capillary flow in soil media, which can be minimized by adaptive consideration of the parameters, such as wall adhesion and appropriate wettability. Full article
Show Figures

Figure 1

Figure 1
<p>The test bed of the green roof with a string wound filter in the middle: (<b>a</b>) test bed; (<b>b</b>) irrigation system.</p>
Full article ">Figure 2
<p>Model of the 3D geometry of the test bed.</p>
Full article ">Figure 3
<p>Model of the 2D symmetrical geometry of the test bed.</p>
Full article ">Figure 4
<p>The generated mesh for the test bed.</p>
Full article ">Figure 5
<p>Simulation and experimental hydrographs.</p>
Full article ">Figure 6
<p>The volume fraction of water in a 2D plane: (<b>a</b>) 53 min and (<b>b</b>) 85 min.</p>
Full article ">
13 pages, 5792 KiB  
Article
Modification of the Surface Crystallinity of Polyphenylene Sulfide and Polyphthalamide Treated by a Pulsed-Arc Atmospheric Pressure Plasma Jet
by Abdessadk Anagri, Sarab Ben Saïd, Cyrille Bazin, Farzaneh Arefi-Khonsari and Jerome Pulpytel
Polymers 2024, 16(18), 2582; https://doi.org/10.3390/polym16182582 - 12 Sep 2024
Viewed by 294
Abstract
Atmospheric plasma jets generated from air or nitrogen using commercial sources with relatively high energy densities are commonly used for industrial applications related to surface treatments, especially to increase the wettability of polymers or to deposit thin films. The heat fluxes to which [...] Read more.
Atmospheric plasma jets generated from air or nitrogen using commercial sources with relatively high energy densities are commonly used for industrial applications related to surface treatments, especially to increase the wettability of polymers or to deposit thin films. The heat fluxes to which the substrates are subjected are typically in the order of 100–300 W/cm2, depending on the treatment conditions. The temperature rise in the treated polymer substrates can have critical consequences, such as a change in the surface crystallinity or even the surface degradation of the materials. In this work, we report the phase transitions of two semicrystalline industrial-grade polymer resins reinforced with glass fibers, namely polyphenylene sulfide (PPS) and polyphthalamide (PPA), subjected to plasma treatments, as well as the modeling of the associated heat transfer phenomena using COMSOL Multiphysics. Depending on the treatment time, the surface of PPS becomes more amorphous, while PPA becomes more crystalline. These results show that the thermal history of the materials must be considered when implementing surface engineering by this type of plasma discharge. Full article
(This article belongs to the Special Issue Plasma Processing of Polymers, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photographs of (<b>a</b>) the free plasma jet and (<b>b</b>) the plasma jet impinging on a substrate; (<b>c</b>) flow regions of an impinging jet (Adapted with permission from [<a href="#B17-polymers-16-02582" class="html-bibr">17</a>], Elsevier, 2006).</p>
Full article ">Figure 2
<p>Evolution of the thermocouple temperature as a function of (<b>a</b>) <span class="html-italic">T<sub>jet</sub></span> with <span class="html-italic">h</span> = 2000 W/m<sup>2</sup>.K and (<b>b</b>) <span class="html-italic">h</span> with <span class="html-italic">T<sub>jet</sub></span> = 793 K.</p>
Full article ">Figure 3
<p>Substrate meshing, plasma torch trajectory, and heat transfer mechanisms.</p>
Full article ">Figure 4
<p>The surface temperature of PPS at <span class="html-italic">t</span> = 9 s. The first 4 lines of the torch pattern are also shown.</p>
Full article ">Figure 5
<p>The maximum temperature on each line as a function of the line speed.</p>
Full article ">Figure 6
<p>The temperature profile on a line as a function of time. The maximum temperature observed on each line corresponds to the center of the plasma jet. The line speed was 5 m/min.</p>
Full article ">Figure 7
<p>(<b>a</b>) Temperature profile along the depth of the material at a given time and (<b>b</b>) along the cut line at the same time. The line speed was 5 m/min.</p>
Full article ">Figure 8
<p>(<b>a</b>) Deconvolution of the WAXD diffractograms of untreated PPS using 3 Gaussian components at 2θ = 18.6°, 19.9° and 20.5°. The baseline was corrected by interpolation and <span class="html-italic">R</span><sup>2</sup> = 0.994; (<b>b</b>) diffractograms of PPS as a function of line speed.</p>
Full article ">Figure 9
<p>(<b>a</b>) Deconvolution of the WAXD diffractograms of untreated PPA using 5 Gaussian components at 2θ = 18.4° and 22.4°, assigned to α1 and α2 crystal, respectively, and 20.3° assigned to the γ crystal phase at 2θ = 20.2 and 24.9, assigned to the amorphous phase and glass fiber, respectively. The baseline was corrected by interpolation and <span class="html-italic">R</span><sup>2</sup> = 0.997; (<b>b</b>) diffractograms of PPA as a function of line speed.</p>
Full article ">Figure 10
<p>Evolution of the percentage of amorphous phase in PPS and PPA treated by air or nitrogen plasma as a function of line speed.</p>
Full article ">Figure 11
<p>DSC curve at a heating rate of 10 K/min for PPA composite; the first and second heat scans.</p>
Full article ">Figure 12
<p>DSC curve at a heating rate of 10 K/min for PPS composite; the first and second heat scans.</p>
Full article ">
13 pages, 12554 KiB  
Article
Wettability Behaviour of Metal Surfaces after Sequential Nanosecond and Picosecond Laser Texturing
by Yin Tang, Zheng Fang, Yang Fei, Shuai Wang, Walter Perrie, Stuart Edwardson and Geoff Dearden
Micromachines 2024, 15(9), 1146; https://doi.org/10.3390/mi15091146 - 12 Sep 2024
Viewed by 344
Abstract
This study examines the wettability behaviour of 304 stainless steel (304SS) and Ti-6Al-4V (Ti64) surfaces after sequential nanosecond (ns) and picosecond (ps) laser texturing; in particular, how the multi-scale surface structures created influence the lifecycle of surface hydrophobicity. The effect of different post-process [...] Read more.
This study examines the wettability behaviour of 304 stainless steel (304SS) and Ti-6Al-4V (Ti64) surfaces after sequential nanosecond (ns) and picosecond (ps) laser texturing; in particular, how the multi-scale surface structures created influence the lifecycle of surface hydrophobicity. The effect of different post-process treatments is also examined. Surfaces were analysed using Scanning Electron Microscopy (SEM), a white light interferometer optical profiler, and Energy Dispersive X-ray (EDX) spectroscopy. Wettability was assessed through sessile drop contact angle (CA) measurements, conducted at regular intervals over periods of up to 12 months, while EDX scans monitored elemental chemical changes. The results show that sequential (ns + ps) laser processing produced multi-scale surface texture with laser-induced periodic surface structures (LIPSS). Compared to the ns laser case, the (ns + ps) laser processed surfaces transitioned more rapidly to a hydrophobic state and maintained this property for much longer, especially when the single post-process treatment was ultrasonic cleaning. Some interesting features in CA development over these extended timescales are revealed. For 304SS, hydrophobicity was reached in 1–2 days, with the CA then remaining in the range of 120 to 140° for up to 180 days; whereas the ns laser-processed surfaces took longer to reach hydrophobicity and only maintained the condition for up to 30 days. Similar results were found for the case of Ti64. The findings show that such multi-scale structured metal surfaces can offer relatively stable hydrophobic properties, the lifetime of which can be extended significantly through the appropriate selection of laser process parameters and post-process treatment. The addition of LIPSS appears to help extend the longevity of the hydrophobic property. In seeking to identify other factors influencing wettability, from our EDX results, we observed a significant and steady rate of increase in the carbon content at the surface over the study period. Full article
(This article belongs to the Special Issue Ultrafast Laser Micro- and Nanoprocessing, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Schematic of optical setup used for ps laser processing (Created with BioRender.com). The beam was attenuated by a λ/2 plate and a Glan-Laser Calcite Polarizer transmitting horizontal polarization. The beam passes through a diffraction-limited beam expander (Rodenstock; M = x3) and is then modulated by a reflective phase-only SLM and input to a galvo scanner after passing through a 4f optical system. An aperture allows the shaped zero-order light to pass through for laser processing.</p>
Full article ">Figure 2
<p>Schematic of the experimental process: laser process and surface property analysis methods (Created with BioRender.com). Four groups (1–4) of polished samples were processed with the ns laser to form the underlying micron-scale doubly periodic structure, while groups 3 and 4 were then also exposed to ps laser processing for LIPSS overlaying. For comparison, groups 1 and 3 were subjected to wettability tests twice a week, while groups 2 and 4 were measured for elemental concentration by EDX once a week. All samples were stored in ambient air during ageing.</p>
Full article ">Figure 3
<p>(Answer for question 2, reviewer 1). SEM images of ns laser processed 304SS and Ti64 surfaces. (<b>a</b>,<b>c</b>) are ×150 magnification, while (<b>b</b>,<b>d</b>) are ×1000 magnification.</p>
Full article ">Figure 4
<p>SEM image of LIPSS on 304SS and Ti64 surfaces: (<b>a</b>,<b>c</b>) ×1000 magnification and (<b>b</b>,<b>d</b>) ×5000 magnification.</p>
Full article ">Figure 5
<p>Surface topography 2D analyses, microscope images, and topography Fourier transform analyses of laser-textured functional surfaces before and after adding ps laser-generated LIPSS.</p>
Full article ">Figure 6
<p>Long-term wettability behaviour of ns laser ablated 304SS surfaces, for different post-process treatments: CA measured during the period up to 257 days after exposure.</p>
Full article ">Figure 7
<p>Wettability behaviour of ns laser processed Ti64 surfaces over a period of 141 days.</p>
Full article ">Figure 8
<p>Wettability behaviour of ns + ps laser processed 304SS and Ti64 surfaces over a period of 141 days.</p>
Full article ">Figure 9
<p>Comparison of time-dependent variation in carbon composition at the surface of samples processed by ns and (ns + ps) lasers, followed by ultrasonic cleaning: (<b>a</b>) 304SS and (<b>b</b>) Ti64.</p>
Full article ">Figure 10
<p>Comparison of time-dependent variation in oxygen composition at the surface of samples processed by ns and (ns + ps) lasers, followed by ultrasonic cleaning: (<b>a</b>) 304SS and (<b>b</b>) Ti64.</p>
Full article ">Figure 11
<p>EDX element-layered distribution map on ns (<b>a</b>) 304SS and (<b>d</b>) Ti64 and ns + ps (<b>b</b>) 304SS and (<b>e</b>) Ti64 laser processed surface structure ×500 magnification and (<b>c</b>) 304SS and (<b>f</b>) Ti64 ×1000 magnification Ns laser processed structure has obvious wider and stronger carbon absorption and oxidises around the peak of the structure compared to the ns + ps laser processed structure.</p>
Full article ">Figure 12
<p>SEM images of 304SS on (<b>a</b>) day 2 and (<b>b</b>) day 142. ns (<b>a-1</b>,<b>b-1</b>) and ns + ps (<b>a-2</b>,<b>b-2</b>) laser processed surface structure ×1000 magnification and (<b>a-3</b>,<b>b-3</b>) ×5000 magnification.</p>
Full article ">Figure 13
<p>SEM images of Ti64 on (<b>a</b>) day 2 and (<b>b</b>) day 142. ns (<b>a-1</b>,<b>b-1</b>) and ns + ps (<b>a-2</b>,<b>b-2</b>) laser-processed surface structure ×1000 magnification and (<b>a-3</b>,<b>b-3</b>) ×5000 magnification.</p>
Full article ">
22 pages, 6976 KiB  
Article
Comparison of Key Properties of Ag-TiO2 and Hydroxyapatite-Ag-TiO2 Coatings on NiTi SMA
by Karolina Dudek, Aleksandra Strach, Daniel Wasilkowski, Bożena Łosiewicz, Julian Kubisztal, Anna Mrozek-Wilczkiewicz, Patryk Zioła and Adrian Barylski
J. Funct. Biomater. 2024, 15(9), 264; https://doi.org/10.3390/jfb15090264 - 12 Sep 2024
Viewed by 324
Abstract
To functionalize the NiTi alloy, multifunctional innovative nanocoatings of Ag-TiO2 and Ag-TiO2 doped with hydroxyapatite were engineered on its surface. The coatings were thoroughly characterized, focusing on surface topography and key functional properties, including adhesion, surface wettability, biocompatibility, antibacterial activity, and [...] Read more.
To functionalize the NiTi alloy, multifunctional innovative nanocoatings of Ag-TiO2 and Ag-TiO2 doped with hydroxyapatite were engineered on its surface. The coatings were thoroughly characterized, focusing on surface topography and key functional properties, including adhesion, surface wettability, biocompatibility, antibacterial activity, and corrosion resistance. The electrochemical corrosion kinetics in a simulated body fluid and the mechanisms were analyzed. The coatings exhibited hydrophilic properties and were biocompatible with fibroblast and osteoblast cells while also demonstrating antibacterial activity against E. coli and S. epidermidis. The coatings adhered strongly to the NiTi substrate, with superior adhesion observed in the hydroxyapatite-doped layers. Conversely, the Ag-TiO2 layers showed enhanced corrosion resistance. Full article
(This article belongs to the Special Issue Advances in Biomedical Alloys and Surface Modification)
Show Figures

Figure 1

Figure 1
<p>Schematic presentation of experiments, including non-exposed bacterial culture (control) (C), NiTi substrate with <span class="html-italic">E. coli</span> (1), HAp coating with <span class="html-italic">E. coli</span> (2), Ag-TiO<sub>2</sub> coating with <span class="html-italic">E. coli</span> (3), and HAp-Ag-TiO<sub>2</sub> coating with <span class="html-italic">E. coli</span> (4). Created in <a href="http://biorinder.com" target="_blank">biorinder.com</a>.</p>
Full article ">Figure 2
<p>SE images (<b>a</b>,<b>c</b>,<b>d</b>,<b>f</b>) and BSE images (<b>b</b>,<b>e</b>) of Ag-TiO<sub>2</sub> (<b>a</b>–<b>c</b>) and HAp-Ag-TiO<sub>2</sub> (<b>d</b>–<b>f</b>) coatings after scratch test with marked critical load (Lc<sub>1</sub>–Lc<sub>2</sub>).</p>
Full article ">Figure 3
<p>AFM topography of the Ag-TiO<sub>2</sub> (<b>a</b>–<b>c</b>) and HAp-Ag-TiO<sub>2</sub> (<b>d</b>–<b>f</b>) coatings.</p>
Full article ">Figure 4
<p>Image of a drop of water on the surface of the Ag-TiO<sub>2</sub> (<b>a</b>) and the HAp-Ag-TiO<sub>2</sub> coating (<b>b</b>).</p>
Full article ">Figure 5
<p>Cytotoxicity of the Ag-TiO<sub>2</sub> (<b>a</b>) and HAp-Ag-TiO<sub>2</sub> (<b>b</b>) against fibroblast (NHDF) and osteoblast (HOB) cells. Microscope imaging of A—NHDF cell line grown on the coatings, B—NHDF cell line grown on the NiTi substrate, C—HOB cell line grown on the coatings, and D—HOB cell line grown on the NiTi substrate for Ag-TiO<sub>2</sub> (<b>c</b>) and HAp-Ag-TiO<sub>2</sub> (<b>d</b>) coatings.</p>
Full article ">Figure 6
<p>Log CFU mL<sup>−1</sup> value for <span class="html-italic">E. coli</span> incubated in Ringer’s solution (<b>A</b>) in culture medium (<b>B</b>), and for <span class="html-italic">S. epidermidis</span> in Ringer’s solution (<b>C</b>) in culture medium (<b>D</b>). Mean + SD (n = 3) with a marked level of significance (* <span class="html-italic">p</span> &lt; 0.05 and *** <span class="html-italic">p</span> &lt; 0.005).</p>
Full article ">Figure 7
<p>Log CFU mL<sup>−1</sup> value for <span class="html-italic">E. coli</span> incubated in Ringer’s solution (<b>A</b>), in culture medium (<b>B</b>), and for <span class="html-italic">S. epidermidis</span> in Ringer’s solution (<b>C</b>), in culture medium (<b>D</b>). Mean +SD (n = 3) with a marked level of significance (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 8
<p>Open circuit potential (E<sub>OC</sub>) as a function of immersion time (t) in the Ringer’s solution at 37 °C for Ag-TiO<sub>2</sub> (<b>a</b>) and HAp-Ag-TiO<sub>2</sub> (<b>b</b>) electrodes.</p>
Full article ">Figure 9
<p>Experimental (symbols) and simulated (CNLS-fit) Bode diagrams as a function of the logarithm of the frequency. (<b>a</b>) Logarithm of the impedance modulus for Ag-TiO<sub>2</sub> electrode. (<b>b</b>) Logarithm of the impedance modulus for HAp-Ag-TiO<sub>2</sub> electrode. (<b>c</b>) Phase angle shift for Ag-TiO<sub>2</sub> electrode; (<b>d</b>) Phase angle shift for HAp-Ag-TiO<sub>2</sub> electrode.</p>
Full article ">Figure 10
<p>Electrical equivalent circuit for pitting corrosion used to model experimental EIS data for electrodes in Ringer’s solution at 37 °C.</p>
Full article ">Figure 11
<p>Cyclic potentiodynamic polarization curve at the polarization scan rate of v = 1 mV s<sup>−1</sup> for the electrode in the Ringer’s solution at 37 °C for Ag-TiO<sub>2</sub> (<b>a</b>) and HAp-Ag-TiO<sub>2</sub> (<b>b</b>) coatings. The arrows indicate the polarization scan direction.</p>
Full article ">Figure 12
<p>CPD surface distribution map (<b>a</b>,<b>c</b>) and CPD histogram with Gaussian fit curve (<b>b</b>,<b>d</b>) for Ag-TiO<sub>2</sub> (<b>a</b>,<b>b</b>) and HAp-Ag-TiO<sub>2</sub> (<b>c</b>,<b>d</b>) coatings.</p>
Full article ">
13 pages, 12300 KiB  
Article
Preparation and Properties of Lightweight Amphiphobic Proppant for Hydraulic Fracturing
by Guang Wang, Qinyue Ma, Longqiang Ren and Jirui Hou
Polymers 2024, 16(18), 2575; https://doi.org/10.3390/polym16182575 - 12 Sep 2024
Viewed by 227
Abstract
The wettability of the proppant is crucial in optimizing the flowback of fracturing fluids and improving the recovery of the produced hydrocarbons. Neutral wet proppants have been proven to improve the fluid flow by reducing the interaction between the fluid and the proppant [...] Read more.
The wettability of the proppant is crucial in optimizing the flowback of fracturing fluids and improving the recovery of the produced hydrocarbons. Neutral wet proppants have been proven to improve the fluid flow by reducing the interaction between the fluid and the proppant surface. In this study, a lightweight amphiphobic proppant (LWAP) was prepared by coating a lightweight ceramic proppant (LWCP) with phenolic resin, epoxy resin, polytetrafluoroethylene (PTFE), and trimethoxy(1H,1H,2H,2H-heptadecafluorodecyl)silane (TMHFS) using a layer-by-layer method. The results indicated that the LWAP exhibited a breakage ratio of 2% under 52 MPa (7.5 K) closure stress, with an apparent density of 2.12 g/cm3 and a bulk density of 1.21 g/cm3. The contact angles of water and olive oil were 125° and 104°, respectively, changing to 124° and 96° after displacement by water and diesel oil. A comparison showed that the LWAP could transport over a significantly longer distance than the LWCP, with the length increasing by more than 80%. Meanwhile, the LWAP displayed notable resistance to scale deposition on the proppant surface compared to the LWCP. Furthermore, the maintained conductivity of the LWAP was higher than that of the LWCP after displacement by water and oil phases alternately. The modified proppant could minimize production declines during hydrocarbon extraction in unconventional reservoirs. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematics of the modification process.</p>
Full article ">Figure 2
<p>(<b>a</b>) Photograph of fracture model apparatus; (<b>b</b>) schematic of plexiglass fracture cell (not to scale).</p>
Full article ">Figure 3
<p>Optical and SEM images of proppants: (<b>a</b>,<b>c</b>) LWCP; (<b>b</b>,<b>d</b>) LWAP.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM images of various scales of LWAP grains; (<b>b</b>) images of liquid drops on LWAP; (<b>c</b>) images of water drop on LWAP; (<b>d</b>) images of olive oil drop on LWAP.</p>
Full article ">Figure 5
<p>Images of LWAP suspended on the surfaces of different liquids; (<b>a</b>) LWAP suspended on the surfaces of water; (<b>b</b>) LWAP suspended on the surfaces of silicone oil.</p>
Full article ">Figure 6
<p>Liquid drops on LWAP under displacement with water and oil; (<b>a</b>) images of liquid drops on LWAP; (<b>b</b>) images of water drop on LWAP; (<b>c</b>) images of olive oil drop on LWAP.</p>
Full article ">Figure 7
<p>Images of proppant placement in fracture systems and the state in water after stirring (2 ppm sodium dodecyl sulfate solution): (<b>a</b>) LWCP placement in fracture system; (<b>b</b>) LWAP placement in fracture system; (<b>c</b>,<b>d</b>) the state in water after stirring. (The blue lines in (<b>a</b>,<b>b</b>) are auxiliary curves representing the settled proppant bed heights; The red lines in (<b>a</b>,<b>b</b>) are auxiliary curves representing the settled proppant bed heights; The area between the right and middle red lines (<b>a</b>,<b>b</b>)represents the length of the LWCP settling bed, and the area between the right and left red lines (<b>a</b>,<b>b</b>) represents the length of the LWAP settling bed.).</p>
Full article ">Figure 8
<p>SEM images and EDX of proppant: LWCP (<b>a</b>,<b>b</b>); LWCP immersed in brine (<b>c</b>,<b>d</b>); LWAP immersed in brine (<b>e</b>,<b>f</b>). (The red boxes in (<b>a</b>,<b>c</b>,<b>e</b>) are the EDX spectral scanning sampling areas).</p>
Full article ">Figure 9
<p>Conductivity (<b>a</b>) and maintained conductivity under 40 MPa (<b>b</b>) of different proppants using DI water as a flowing fluid (LWCP1 and LWAP1 after water and oil displacement).</p>
Full article ">
12 pages, 7370 KiB  
Article
Impact of Surface Pretreatment on the Corrosion Resistance and Adhesion of Thin Film Coating on SS316L Bipolar Plates for Proton-Exchange Membrane Fuel Cell Applications
by Yasin Mehdizadeh Chellehbari, Abhay Gupta, Xianguo Li and Samaneh Shahgaldi
Molecules 2024, 29(18), 4319; https://doi.org/10.3390/molecules29184319 - 12 Sep 2024
Viewed by 359
Abstract
Coated SS316L is a potential alternative to the graphite bipolar plates (BPPs) used in proton-exchange membrane fuel cells (PEMFCs) owing to their low manufacturing cost and machinability. Due to their susceptibility to corrosion and passivation, which increases PEMFC ohmic resistance, protective and conductive [...] Read more.
Coated SS316L is a potential alternative to the graphite bipolar plates (BPPs) used in proton-exchange membrane fuel cells (PEMFCs) owing to their low manufacturing cost and machinability. Due to their susceptibility to corrosion and passivation, which increases PEMFC ohmic resistance, protective and conductive coatings on SS316L have been developed. However, coating adhesion is one of the challenges in the harsh acidic environment of PEMFCs, affecting the performance and durability of BPPs. This study compares mechanical polishing and the frequently adopted chemical etchants for SS316L: Adler’s, V2A, and Carpenter’s etchant with different etching durations and their impact on the wettability, adhesion, and corrosion resistance of a Nb-coated SS316L substrate. Contact angle measurements and laser microscopy revealed that all etching treatments increased the hydrophobicity and surface roughness of SS316L substrates. Ex situ potentiodynamic and potentiostatic polarization tests and interfacial contact resistance analysis revealed high corrosion resistance, interfacial conductivity, and adhesion of the Nb-coated SS316L substrate pretreated with V2A (7 min) and Adler’s (3 min) etchant. Increased hydrophobicity (contact angle = 101°) and surface roughness (Ra = 74 nm) achieved using V2A etchant led to the lowest corrosion rate (3.3 µA.cm−2) and interfacial resistance (15.4 mΩ.cm2). This study established pretreatment with V2A etchant (a solution of HNO3, HCl, and DI water (1:9:23 mole ratio)) as a promising approach for improving the longevity, electrochemical stability, and efficiency of the coated SS316L BPPs for PEMFC application. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of the arithmetic mean roughness (Ra) and arithmetic mean height (Sa), line roughness profile, and morphology of the SS316L surfaces with different pretreatments.</p>
Full article ">Figure 2
<p>Comparison of surface morphology, mean surface roughness (Ra) and mean surface height (Sa), contact angle, and adhesion impact of the SS316L surfaces with different pretreatments.</p>
Full article ">Figure 3
<p>(<b>a</b>) Comparison of surface morphology for all coated samples. (<b>b</b>) X-ray diffractograms of uncoated and niobium (Nb)-coated SS316L samples.</p>
Full article ">Figure 4
<p>Tafel plots recorded before potentiostatic polarization in a 0.5 M H<sub>2</sub>SO<sub>4</sub> solution at 70 °C for all samples.</p>
Full article ">Figure 5
<p>The figure shows 6 h potentiostatic polarization at +0.8 V vs. reversible hydrogen electrode (RHE) in 0.5 M H<sub>2</sub>SO<sub>4</sub>, maintained at 70 °C for all samples.</p>
Full article ">Figure 6
<p>Interfacial contact resistance (ICR) plots (<b>a</b>) before corrosion analysis, (<b>b</b>) after corrosion analysis, and (<b>c</b>) at the compaction force of 1.5 MPa.</p>
Full article ">
Back to TopTop