[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
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
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
remove_circle_outline
remove_circle_outline

Search Results (38,787)

Search Parameters:
Keywords = polymers

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 11742 KiB  
Article
Large Enhancement of Photoluminescence Obtained in Thin Polyfluorene Films of Optimized Microstructure
by Otto Todor-Boer, Cosmin Farcău and Ioan Botiz
Polymers 2024, 16(16), 2278; https://doi.org/10.3390/polym16162278 (registering DOI) - 11 Aug 2024
Viewed by 84
Abstract
There is a clearly demonstrated relationship between the microstructure, processing and resulting optoelectronic properties of conjugated polymers. Here, we exploited this relationship by exposing polyfluorene thin films to various solvent vapors via confined-solvent vapor annealing to optimize their microstructure, with the final goal [...] Read more.
There is a clearly demonstrated relationship between the microstructure, processing and resulting optoelectronic properties of conjugated polymers. Here, we exploited this relationship by exposing polyfluorene thin films to various solvent vapors via confined-solvent vapor annealing to optimize their microstructure, with the final goal being to enhance their emission properties. Our results have demonstrated enlargements in photoluminescence intensity of up to 270%, 258% and 240% when thin films of polyfluorenes of average molecular weights of 105,491 g/mol, 63,114 g/mol and 14,000 g/mol, respectively, experienced increases in their β-phase fractions upon processing. Full article
(This article belongs to the Section Polymer Membranes and Films)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the C-SVA home-made device. This device is mainly composed of a sample chamber made of aluminum and equipped with a PT100 temperature sensor. Beneath there are placed a Peltier element and a heat sink. This ensemble is connected to a temperature controller, a power source and a gas bubbling system. The latter is able to introduce a regulated amount of solvent vapors into the sample chamber.</p>
Full article ">Figure 2
<p>Chemical structure of PFO.</p>
Full article ">Figure 3
<p>PL (<b>a</b>) and normalized PL (<b>b</b>) spectra of PFO<sub>105k</sub> thin films before (black) and after their processing via the C-SVA method in tetrahydrofuran (blue), chloroform (red) and toluene (olive) vapors, along with their corresponding digital images (<b>c</b>–<b>f</b>), respectively. All PL spectra were acquired using an excitation wavelength of 390 nm.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>l</b>) AFM height (<b>a</b>–<b>d</b>) and phase (<b>e</b>–<b>l</b>) micrographs depicting the morphology of PFO<sub>105k</sub> thin films before (<b>a</b>,<b>e</b>,<b>i</b>) and after their exposure via the C-SVA method to chloroform (<b>b</b>,<b>f</b>,<b>j</b>), tetrahydrofuran (<b>c</b>,<b>g</b>,<b>k</b>) and toluene (<b>d</b>,<b>h</b>,<b>l</b>) vapors.</p>
Full article ">Figure 5
<p>(<b>a</b>) Raman spectra recorded for a spin-cast film of PFO<sub>105k</sub> before (black) and after its processing via the C-SVA method in tetrahydrofuran (blue), chloroform (red) and toluene (olive) vapors, respectively. The inset depicts a zoom-in of the main Raman peak located at around 1604 cm<sup>−1</sup>. (<b>b</b>) The same vertically translated Raman spectra zoomed-in in the 1000–1500 cm<sup>−1</sup> spectral interval, as indicated by the dotted shape in (<b>a</b>), emphasizing the changes of various peaks induced upon C-SVA processing. Grey vertical dotted lines/arrows in (<b>b</b>) are for guiding the eye only.</p>
Full article ">Figure 6
<p>AFM height (<b>a</b>–<b>d</b>) and phase (<b>e</b>–<b>l</b>) micrographs depicting the surface morphology of PFO<sub>63k</sub> thin films before (<b>a</b>,<b>e</b>,<b>i</b>) and after their exposure via the C-SVA method to chloroform (<b>b</b>,<b>f</b>,<b>j</b>), tetrahydrofuran (<b>c</b>,<b>g</b>,<b>k</b>) and toluene (<b>d</b>,<b>h</b>,<b>l</b>) vapors.</p>
Full article ">Figure 7
<p>AFM height (<b>a</b>–<b>d</b>) and phase (<b>e</b>–<b>l</b>) micrographs depicting the surface morphology of PFO<sub>14k</sub> thin films before (<b>a</b>,<b>e</b>,<b>i</b>) and after their exposure via the C-SVA method to chloroform (<b>b</b>,<b>f</b>,<b>j</b>), tetrahydrofuran (<b>c</b>,<b>g</b>,<b>k</b>) and toluene (<b>d</b>,<b>h</b>,<b>l</b>) vapors.</p>
Full article ">
14 pages, 4295 KiB  
Article
Bio-Based Polyurethane–Urea with Self-Healing and Closed-Loop Recyclability Synthesized from Renewable Carbon Dioxide and Vanillin
by Tianyi Han, Tongshuai Tian, Shan Jiang and Bo Lu
Polymers 2024, 16(16), 2277; https://doi.org/10.3390/polym16162277 (registering DOI) - 10 Aug 2024
Viewed by 328
Abstract
Developing recyclable and self-healing non-isocyanate polyurethane (NIPU) from renewable resources to replace traditional petroleum-based polyurethane (PU) is crucial for advancing green chemistry and sustainable development. Herein, a series of innovative cross-linked Poly(hydroxyurethane-urea)s (PHUUs) were prepared using renewable carbon dioxide (CO2) and [...] Read more.
Developing recyclable and self-healing non-isocyanate polyurethane (NIPU) from renewable resources to replace traditional petroleum-based polyurethane (PU) is crucial for advancing green chemistry and sustainable development. Herein, a series of innovative cross-linked Poly(hydroxyurethane-urea)s (PHUUs) were prepared using renewable carbon dioxide (CO2) and vanillin, which displayed excellent thermal stability properties and solvent resistance. These PHUUs were constructed through the introduction of reversible hydrogen and imine bonds into cross-linked polymer networks, resulting in the cross-linked PHUUs exhibiting thermoplastic-like reprocessability, self healing, and closed-loop recyclability. Notably, the results indicated that the VL-TTD*-50 with remarkable hot-pressed remolding efficiency (nearly 98.0%) and self-healing efficiency (exceeding 95.0%) of tensile strength at 60 °C. Furthermore, they can be degraded in the 1M HCl and THF (v:v = 2:8) solution at room temperature, followed by regeneration without altering their original chemical structure and mechanical properties. This study presents a novel strategy for preparing cross-linked PHUUs with self-healing and closed-loop recyclability from renewable resources as sustainable alternatives for traditional petroleum-based PUs. Full article
(This article belongs to the Special Issue Preparation and Application of Biodegradable Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic structure of VL-TTD* elastomers.</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra of VL−C and VL−TTD*s; (<b>b</b>) FTIR spectra of VL−TTD*s; (<b>c</b>) swelling rates of the VL−TTD*s; (<b>d</b>) gel contents of the VL−TTD*s.</p>
Full article ">Figure 3
<p>Temperature-dependent <sup>1</sup>H NMR spectra of VL-TTD*-50 upon heating from 30 to 100 °C (<b>a</b>) and upon cooling from 100 to 30 °C (<b>b</b>).</p>
Full article ">Figure 4
<p>The rheology of VL-TTD*s: (<b>a</b>) Variation in the storage modulus as a function of the frequency and (<b>b</b>) variation in the loss modulus as a function of the frequency.</p>
Full article ">Figure 5
<p>(<b>a</b>) TGA and DTG curves of VL-TTD*s in N<sub>2</sub>; (<b>b</b>) DSC curves of VL-TTD*s; (<b>c</b>) stress–strain curves of VL-TTD*s; (<b>d</b>) mechanical properties of VL-TTD*s.</p>
Full article ">Figure 6
<p>(<b>a</b>) Photographs of the hot-pressing process. (<b>b</b>) FT-IR spectra of VL-TTD*-50 after two recycling processes; (<b>c</b>) stress–strain curves of VL-TTD*-50 after two recycling processes.</p>
Full article ">Figure 7
<p>(<b>a</b>) Polarizing optical microscopy images of the self-healing process of a crack on the VL-TTD*-50; (<b>b</b>) images of the self-healing of VL-TTD*-50: (i) original sample; (ii) cut segments; (iii) healed sample; (iv) the healed sample with the hanging weight of a 500 g bottle; (<b>c</b>) stress–strain curves of the VL-TTD*-50 after healing at different times.</p>
Full article ">Figure 8
<p>The self-healing mechanism of the VL-TTD*s.</p>
Full article ">Figure 9
<p>Closed-loop recycling of VL-TTD*s.</p>
Full article ">Figure 10
<p>(<b>a</b>) Degradation rate curves of VL-TTD*-50 in 1M HCl and THF (<span class="html-italic">v</span>:<span class="html-italic">v</span> = 2:8) and H<sub>2</sub>O/THF (<span class="html-italic">v</span>:<span class="html-italic">v</span> = 2:8); (<b>b</b>) FT-IR spectra of VL-TTD*-50, degraded products, and regenerated VL-TTD*-50; (<b>c</b>) stress–strain curves of original and regenerated VL-TTD*-50; (<b>d</b>) DSC curves of original and regenerated VL-TTD*-50.</p>
Full article ">Scheme 1
<p>Synthetic routes of the VL-TTD*s.</p>
Full article ">
15 pages, 9001 KiB  
Article
Novel Water Probe for High-Frequency Focused Transducer Applied to Scanning Acoustic Microscopy System: Simulation and Experimental Investigation
by Van Hiep Pham, Le Hai Tran, Jaeyeop Choi, Hoanh-Son Truong, Tan Hung Vo, Dinh Dat Vu, Sumin Park and Junghwan Oh
Sensors 2024, 24(16), 5179; https://doi.org/10.3390/s24165179 (registering DOI) - 10 Aug 2024
Viewed by 317
Abstract
A scanning acoustic microscopy (SAM) system is a common non-destructive instrument which is used to evaluate the material quality in scientific and industrial applications. Technically, the tested sample is immersed in water during the scanning process. Therefore, a robot arm is incorporated into [...] Read more.
A scanning acoustic microscopy (SAM) system is a common non-destructive instrument which is used to evaluate the material quality in scientific and industrial applications. Technically, the tested sample is immersed in water during the scanning process. Therefore, a robot arm is incorporated into the SAM system to transfer the sample for in-line inspection, which makes the system complex and increases time consumption. The main aim of this study is to develop a novel water probe for the SAM system, that is, a waterstream. During the scanning process, water was supplied using a waterstream instead of immersing the sample in the water, which leads to a simple design of an automotive SAM system and a reduction in time consumption. In addition, using a waterstream in the SAM system can avoid contamination of the sample due to immersion in water for long-time scanning. Waterstream was designed based on the measured focal length calculation of the transducer and simulated to investigate the internal flow characteristics. To validate the simulation results, the waterstream was prototyped and applied to the TSAM-400 and W-FSAM traditional and fast SAM systems to successfully image some samples such as carbon fiber-reinforced polymers, a printed circuit board, and a 6-inch wafer. These results demonstrate the design method of the water probe applied to the SAM system. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the TSAM−400 system with waterstream.</p>
Full article ">Figure 2
<p>Schematic of the W−FSAM system with waterstream.</p>
Full article ">Figure 3
<p>The measured focal length of transducer inside sample.</p>
Full article ">Figure 4
<p>Water domain modeling.</p>
Full article ">Figure 5
<p>Mesh construction of (<b>a</b>) water domain, (<b>b</b>) inlet, and (<b>c</b>) outlet.</p>
Full article ">Figure 6
<p>Velocity distribution inside water domain plotted as streamlines.</p>
Full article ">Figure 7
<p>Pressure distribution inside water domain plotted as streamlines.</p>
Full article ">Figure 8
<p>The velocity and pressure distribution along the centerline between the transducer and outlet.</p>
Full article ">Figure 9
<p>Waterstream design and prototype: (<b>a</b>) 2D drawing, (<b>b</b>) 3D exploded drawing, (<b>c</b>) rendered concept, (<b>d</b>) prototype.</p>
Full article ">Figure 10
<p>Rendered image of TSAM−400 system with two waterstreams.</p>
Full article ">Figure 11
<p>(<b>a</b>) CFRP sample: C-scan images of (<b>b</b>) top surface, (<b>c</b>) underlayer, and (<b>d</b>) enlarged view of underlayer.</p>
Full article ">Figure 12
<p>(<b>a</b>) PCB sample, (<b>b</b>) top surface C-scan image, (<b>c</b>) underlayer C-scan image, and (<b>d</b>) enlarged view of soldering area.</p>
Full article ">Figure 13
<p>(<b>a</b>) The 6-inch wafer sample, (<b>b</b>) C-scan image of wafer, (<b>c</b>) enlarged view of I area, (<b>d</b>) enlarged view of II area.</p>
Full article ">
18 pages, 8055 KiB  
Review
The Development of Polylactide Nanocomposites: A Review
by Purba Purnama, Zaki Saptari Saldi and Muhammad Samsuri
J. Compos. Sci. 2024, 8(8), 317; https://doi.org/10.3390/jcs8080317 (registering DOI) - 10 Aug 2024
Viewed by 164
Abstract
Polylactide materials present a promising alternative to petroleum-based polymers due to their sustainability and biodegradability, although they have certain limitations in physical and mechanical properties for specific applications. The incorporation of nanoparticles, such as layered silicate (clay), carbon nanotubes, metal or metal oxide, [...] Read more.
Polylactide materials present a promising alternative to petroleum-based polymers due to their sustainability and biodegradability, although they have certain limitations in physical and mechanical properties for specific applications. The incorporation of nanoparticles, such as layered silicate (clay), carbon nanotubes, metal or metal oxide, cellulose nanowhiskers, can address these limitations by enhancing the thermal, mechanicals, barriers, and some other properties of polylactide. However, the distinct characteristics of these nanoparticles can affect the compatibility and processing of polylactide blends. In the polylactide nanocomposites, well-dispersed nanoparticles within the polylactide matrix result in excellent mechanical and thermal properties of the materials. Surface modification is required to improve compatibility and the crystallization process in the blended materials. This article reviews the development of polylactide nanocomposites and their applications. It discusses the general aspect of polylactides and nanomaterials as nanofillers, followed by the discussion of the processing and characterization of polylactide nanocomposites, including their applications. The final section summarizes and discusses the future challenges of polylactide nanocomposites concerning the future material’s requirements and economic considerations. As eco-friendly materials, polylactide nanocomposites offer significant potential to replace petroleum-based polymers. Full article
(This article belongs to the Special Issue Sustainable Biocomposites, Volume II)
13 pages, 2932 KiB  
Article
Snapshot Multi-Wavelength Birefringence Imaging
by Shuang Wang, Xie Han and Kewu Li
Sensors 2024, 24(16), 5174; https://doi.org/10.3390/s24165174 (registering DOI) - 10 Aug 2024
Viewed by 153
Abstract
A snapshot multi-wavelength birefringence imaging measurement method was proposed in this study. The RGB-LEDs at wavelengths 463 nm, 533 nm, and 629 nm were illuminated with circularly polarized light after passing through a circular polarizer. The transmitted light through the birefringent sample was [...] Read more.
A snapshot multi-wavelength birefringence imaging measurement method was proposed in this study. The RGB-LEDs at wavelengths 463 nm, 533 nm, and 629 nm were illuminated with circularly polarized light after passing through a circular polarizer. The transmitted light through the birefringent sample was captured by a color polarization camera. A single imaging process captured light intensity in four polarization directions (0°, 45°, 90°, and 135°) for each of the three RGB spectral wavelength channels, and subsequently measured the first three elements of Stokes vectors (S0, S1, and S2) after the sample. The birefringence retardance and fast-axis azimuthal angle were determined simultaneously. An experimental setup was constructed, and polarization response matrices were calibrated for each spectral wavelength channel to ensure the accurate detection of Stokes vectors. A polymer true zero-order quarter-wave plate was employed to validate measurement accuracy and repeatability. Additionally, stress-induced birefringence in a PMMA arch-shaped workpiece was measured both before and after the application of force. Experimental results revealed that the repeatability of birefringence retardance and fast-axis azimuthal angle was better than 0.67 nm and 0.08°, respectively. This approach enables multispectral wavelength, high-speed, high-precision, and high-repeatability birefringence imaging measurements through a single imaging session. Full article
(This article belongs to the Section Sensing and Imaging)
Show Figures

Figure 1

Figure 1
<p>Birefringence imaging system. Schematic (<b>a</b>) and device (<b>b</b>) diagram.</p>
Full article ">Figure 2
<p>RGB-LED spectrum.</p>
Full article ">Figure 3
<p>Normalized intensity <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi mathvariant="normal">I</mi> <mo stretchy="false">(</mo> <mi>λ</mi> <mo stretchy="false">)</mo> </mrow> </msub> </mrow> </semantics></math> vs. retardance <span class="html-italic">R</span>.</p>
Full article ">Figure 4
<p>Calibration of response matrices: (<b>a</b>) B-, (<b>b</b>) G-, and (<b>c</b>) R-color response matrices.</p>
Full article ">Figure 5
<p>Measured light intensity before and after calibration: (<b>a1</b>) B-, (<b>b1</b>) G-, and (<b>c1</b>) R-color channels. Degree of linear polarization: (<b>a2</b>) B-, (<b>b2</b>) G-, and (<b>c2</b>) R-color channels before calibration and (<b>a3</b>) B-, (<b>b3</b>) G-, and (<b>c3</b>) R-color channels after calibration for 0° direction of polarizer.</p>
Full article ">Figure 6
<p>Results of quarter-wave plate: (<b>a</b>–<b>c</b>) retardance at B-, G-, and R-wavelength channels and (<b>d</b>–<b>f</b>) fast-axis azimuthal angle at B-, G-, and R-wavelength channels.</p>
Full article ">Figure 7
<p>Stress birefringence measurements of PMMA arch-shaped workpiece: (<b>a</b>) workpiece, (<b>b1</b>–<b>b3</b>, <b>c1</b>–<b>c3</b>) retardance, and (<b>b4</b>–<b>b6</b>, <b>c4</b>–<b>c6</b>) fast-axis azimuthal angles at B-, G-, and R-wavelength channels before and after applying force. Point A in <a href="#sensors-24-05174-f007" class="html-fig">Figure 7</a>(a) is a region of interest measuring about 3 × 3 pixels.</p>
Full article ">
17 pages, 4465 KiB  
Article
The Development of Sustainable Polyoxymethylene (POM)-Based Composites by the Introduction of Natural Fillers and Melt Blending with Poly(lactic acid)-PLA
by Anna Soćko and Jacek Andrzejewski
J. Compos. Sci. 2024, 8(8), 315; https://doi.org/10.3390/jcs8080315 (registering DOI) - 10 Aug 2024
Viewed by 171
Abstract
The conducted study was focused on the development of a new type of technical blend reinforced with natural fillers. The study was divided into two parts, where, in the first stage of the research, unmodified POM was reinforced with different types of natural [...] Read more.
The conducted study was focused on the development of a new type of technical blend reinforced with natural fillers. The study was divided into two parts, where, in the first stage of the research, unmodified POM was reinforced with different types of natural fillers: cellulose, wood flour, and husk particles. In order to select the type of filler intended for further modification, the mechanical characteristics were assessed. The 20% wood flour (WF) filler system was selected as the reinforcement. The second stage of research involved the use of a combination of polyoxymethylene POM and poly(lactic acid) PLA. The POM/PLA blend (ratio 50/50%) was modified with an elastomeric compound (EBA) and chain extender as the compatibilized reactive (CE). The microscopic analysis revealed that for the POM/PLA system, the filler–matrix interface is characterized by better wettability, which might suggest higher adhesion. The mechanical performance revealed that for POM/PLA-based composites, the properties were very close to the results for POM-WF composites; however, there is still a significant difference in thermal resistance in favor of POM-based materials. The increase in thermomechanical properties for POM/PLA composites occurs after heat treatment. The increasing crystallinity of the PLA phase allows for a significant increase in the heat deflection temperature (HDT), even above 125 °C. Full article
(This article belongs to the Special Issue Progress in Polymer Composites, Volume III)
Show Figures

Figure 1

Figure 1
<p>The results of static tensile tests and Charpy impact resistance measurements. The results presented on the charts reflect the properties of POM-based composites reinforced with different types of fillers: (<b>A</b>) tensile modulus, (<b>B</b>) tensile strength, (<b>C</b>) elongation at break, and (<b>D</b>) Charpy impact strength.</p>
Full article ">Figure 2
<p>The results of static tensile tests and Charpy impact resistance measurements for POM/PLA blends: (<b>A</b>) tensile modulus; (<b>B</b>) tensile strength; (<b>C</b>) elongation at break; (<b>D</b>) Charpy impact strength. Plots reflect the results for unmodified and wood flour (WF)-reinforced samples.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>) The structure of POM-based composites with different types of fillers; (<b>E</b>) the structure of modified POM/WF composites.</p>
Full article ">Figure 4
<p>The results of the DMTA analysis (storage modulus and tan δ): (<b>A</b>) for unmodified samples; (<b>B</b>) for WF-modified materials. (<b>C</b>) The results of the head deflection temperature (HDT) measurements.</p>
Full article ">Figure 5
<p>The results of the DSC analysis for 20% WF reinforced composites. Both charts present the first heating scans: (<b>A</b>) for untreated samples; (<b>B</b>) for annealed composite materials.</p>
Full article ">
28 pages, 3497 KiB  
Review
Polymer-Assisted Graphite Exfoliation: Advancing Nanostructure Preparation and Multifunctional Composites
by Jaime Orellana, Esteban Araya-Hermosilla, Andrea Pucci and Rodrigo Araya-Hermosilla
Polymers 2024, 16(16), 2273; https://doi.org/10.3390/polym16162273 (registering DOI) - 10 Aug 2024
Viewed by 374
Abstract
Exfoliated graphite (ExG) embedded in a polymeric matrix represents an accessible, cost-effective, and sustainable method for generating nanosized graphite-based polymer composites with multifunctional properties. This review article analyzes diverse methods currently used to exfoliate graphite into graphite nanoplatelets, few-layer graphene, and polymer-assisted graphene. [...] Read more.
Exfoliated graphite (ExG) embedded in a polymeric matrix represents an accessible, cost-effective, and sustainable method for generating nanosized graphite-based polymer composites with multifunctional properties. This review article analyzes diverse methods currently used to exfoliate graphite into graphite nanoplatelets, few-layer graphene, and polymer-assisted graphene. It also explores engineered methods for small-scale pilot production of polymer nanocomposites. It highlights the chemistry involved during the graphite intercalation and exfoliation process, particularly emphasizing the interfacial interactions related to steric repulsion forces, van der Waals forces, hydrogen bonds, π-π stacking, and covalent bonds. These interactions promote the dispersion and stabilization of the graphite derivative structures in polymeric matrices. Finally, it compares the enhanced properties of nanocomposites, such as increased thermal and electrical conductivity and electromagnetic interference (EMI) shielding applications, with those of neat polymer materials. Full article
(This article belongs to the Special Issue Functional Graphene-Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Graphite structures and sources.</p>
Full article ">Figure 2
<p>SEM micrographs of rolling intercalation PS with 5% of colloidal graphite [<a href="#B27-polymers-16-02273" class="html-bibr">27</a>]. Reproduced with permission from Tu, H.; Polymers for advanced technologies; Published by John/Wiley &amp; Sons Ltd.; 2008.</p>
Full article ">Figure 3
<p>Orientation of the graphite platelet structures in extruded strands: (left) schematic figure of platelet orientation along strand flow direction by extrusion out of the die; and (right) transmission light microscopy pictures of samples cut perpendicular to the strand direction (shows mainly the layer thickness) and cut parallel to the long-axis of the strand (shows the lateral dimension of visible GNP structures), here shown for 1 wt% Graphene nanopowder AO-3 in PC [<a href="#B124-polymers-16-02273" class="html-bibr">124</a>]. Reproduced with permission from Pötschke, P.; Materials; Published by MDPI; 2017.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>) PVDF/expanded graphite composite fabricated without water (P-EG) and (<b>b</b>) PVDF/expanded graphite composite fabricated with water (P-EG-W) samples [<a href="#B91-polymers-16-02273" class="html-bibr">91</a>]. Reproduced with permission from Tong, J.; Macromolecular materials and engineering; Published by Wiley-VCH Verlag GMBH &amp; Co.; 2020.</p>
Full article ">Figure 5
<p>FESEM images of (<b>A</b>) PCGF-30 [<a href="#B87-polymers-16-02273" class="html-bibr">87</a>]. Reproduced with permission from Pradhan, S.S.; Polymer composites; Published by John/Wiley &amp; Sons Ltd.; 2021.</p>
Full article ">Figure 6
<p>SEM images of graphite at magnifications of 15 and 150 Kx at different ball milling times: (<b>a</b>,<b>b</b>) 0 h; (<b>c</b>,<b>d</b>) 1 h; (<b>e</b>,<b>f</b>) 4 h; (<b>g</b>,<b>h</b>) 8 h; (<b>i</b>,<b>j</b>) 16 h [<a href="#B130-polymers-16-02273" class="html-bibr">130</a>]. Reproduced with permission from Visco, A.; Polymers; Published by MDPI; 2021.</p>
Full article ">Figure 7
<p>EMI-S effectiveness (dB) and wt% filler in different matrix.</p>
Full article ">Figure 8
<p>Thermal conductivity and wt% filler in different matrix, conductive nanocomposite.</p>
Full article ">Figure 9
<p>Thermal conductivity and wt% filler in different matrices, conductive nanocomposite.</p>
Full article ">Scheme 1
<p>Engineering state-of-the-art techniques to generate exfoliated graphite/polymers composites.</p>
Full article ">
22 pages, 2822 KiB  
Article
Automatic Foreign Matter Segmentation System for Superabsorbent Polymer Powder: Application of Diffusion Adversarial Representation Learning
by Ssu-Han Chen, Meng-Jey Youh, Yan-Ru Chen, Jer-Huan Jang, Hung-Yi Chen, Hoang-Giang Cao, Yang-Shen Hsueh, Chuan-Fu Liu and Kevin Fong-Rey Liu
Mathematics 2024, 12(16), 2473; https://doi.org/10.3390/math12162473 (registering DOI) - 10 Aug 2024
Viewed by 177
Abstract
In current industries, sampling inspections of the quality of powders, such as superabsorbent polymers (SAPs) still are conducted via visual inspection. The size of samples and foreign matter are around 500 μm, making them difficult for humans to identify. An automatic foreign matter [...] Read more.
In current industries, sampling inspections of the quality of powders, such as superabsorbent polymers (SAPs) still are conducted via visual inspection. The size of samples and foreign matter are around 500 μm, making them difficult for humans to identify. An automatic foreign matter detection system for powder has been developed in the present study. The powder samples can be automatically delivered, distributed, and recycled, and images of them are captured through the hardware of the system, while the identification software of this system was developed based on diffusion adversarial representation learning (DARL). The background image is a foreign-matter-free powder image with an input image size of 1024 × 1024 × 3. Since DARL includes adversarial segmentation, a diffusion process, and synthetic image generation, the DARL model was trained using a diffusion block with the employment of a U-Net attention mechanism and a spatial-adaptation de-normalization (SPADE) layer through the adoption of a loss function from a vanilla generative adversarial network (GAN). This model was then compared with supervised models such as a fully convolutional network (FCN), U-Net, and DeepLABV3+, as well as with an unsupervised Otsu threshold segmentation. It should be noted that only 10% of the training samples were utilized for the DARL to learn and the intersection over union (IoU) of the DARL can reach up to 80.15%, which is much higher than the 59.00%, 53.47%, 49.39%, and 30.08% for the Otsu threshold segmentation, FCN, U-Net, and DeepLABV3+ models. Therefore, the performance of the model developed in the present study would not be degraded due to an insufficient number of samples containing foreign matter. In practical applications, there is no need to collect, label, and design features for a large number of foreign matter samples before using the developed system. Full article
(This article belongs to the Section Engineering Mathematics)
14 pages, 4836 KiB  
Article
Sustainable Algae-Derived Carbon Particles from Hydrothermal Liquefaction: An Innovative Reinforcing Agent for Epoxy Matrix Composite
by Abhijeet Mali, Philip Agbo, Shobha Mantripragada, Vishwas S. Jadhav, Lijun Wang and Lifeng Zhang
Sustainability 2024, 16(16), 6870; https://doi.org/10.3390/su16166870 (registering DOI) - 10 Aug 2024
Viewed by 236
Abstract
Algae is a promising sustainable feedstock for the generation of bio-crude oil, which is a sustainable alternative to fossil fuels, through the thermochemical process of hydrothermal liquefaction (HTL). However, this process also generates carbon particles (algae-derived carbon, ADC) as a significant byproduct. Herein, [...] Read more.
Algae is a promising sustainable feedstock for the generation of bio-crude oil, which is a sustainable alternative to fossil fuels, through the thermochemical process of hydrothermal liquefaction (HTL). However, this process also generates carbon particles (algae-derived carbon, ADC) as a significant byproduct. Herein, we report a brand-new and value-added use of ADC particles as a reinforcing agent for epoxy matrix composites (EMCs). ADC particles were synthesized through HTL processing of Chlorella vulgaris (a green microalgae) and characterized for morphology, average size, specific surface area, porosity, and functional groups. The ADC particles were subsequently integrated into a representative epoxy resin (EPON 862) as a reinforcing filler at loading levels of 0.25%, 0.5%, 1%, and 2% by weight. The tensile, flexural, and Izod impact properties, as well as the thermal stability, of the resulting EMCs were evaluated. It is revealed that the ADC particles are a sustainable and effective reinforcing agent for EMCs at ultra-low loading. Specifically, the ADC-reinforced EMC with 1 wt.% ADC showed improvements of ~24%, ~30%, ~31%, and ~57% in tensile strength, Young’s modulus, elongation at break, and work of fracture (WOF), respectively, and improvements of ~10%, ~37%, ~24%, and ~39% in flexural strength, flexural modulus, flexural elongation at break, and flexural WOF, respectively, as well as an improvement of ~54% in Izod impact strength, compared to those corresponding properties of neat epoxy. In the meantime, the thermal decomposition temperatures at 60% and 80% weight loss of the abovementioned ADC-reinforced EMC increased from 410 °C to 415 °C and from 448 °C to 515 °C in comparison with those of neat epoxy. This study highlighted the potential of sustainable ADC particles as a reinforcing agent in the field of polymer matrix composite materials, which represented a novel and sustainable approach that would mitigate greenhouse gas remission and reduce reliance on nonrenewable reinforcing fillers in the polymer composite industry. Full article
(This article belongs to the Section Sustainable Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for ADC particle preparation.</p>
Full article ">Figure 2
<p>SEM images of raw ADC particles at low (<b>A</b>) and high (<b>B</b>) magnifications and ADC particles after ultrasonication (<b>C</b>).</p>
Full article ">Figure 3
<p>FTIR (<b>A</b>) and Raman (<b>B</b>) spectra of ADC particles.</p>
Full article ">Figure 4
<p>Full survey (<b>A</b>) and high-resolution C1s (<b>B</b>) XPS spectra of ADC particles.</p>
Full article ">Figure 5
<p>Tensile strength (<b>A</b>), Young’s modulus (<b>B</b>), elongation at break (<b>C</b>), and work of fracture (<b>D</b>) of neat epoxy resin (control sample) and ADC-reinforced EMCs with ADC particles at loadings of 0.25 wt.%, 0.5 wt.%, 1 wt.%, and 2 wt.% from tensile test.</p>
Full article ">Figure 6
<p>Flexural strength (<b>A</b>), flexural modulus (<b>B</b>), flexural elongation at break (<b>C</b>), and flexural work of fracture (<b>D</b>) of neat epoxy resin (control sample) and ADC-reinforced EMCs with ADC particle loadings at 0.25 wt.%, 0.5 wt.%, 1 wt.%, and 2 wt.% from the flexural (three-point bending) test.</p>
Full article ">Figure 7
<p>Izod impact strength of neat epoxy resin (control sample) and ADC-reinforced EMCs with ADC particle loadings at 0.25 wt.%, 0.5 wt.%, 1 wt.%, and 2 wt.% from the Izod impact test.</p>
Full article ">Figure 8
<p>Representative SEM images of fracture surfaces of neat epoxy (<b>A</b>) and the ADC-reinforced EMC with ADC loading of 1 wt.% (<b>B</b>) after mechanical tests. Labels “1”, “2”, and “3” stand for tensile test, flexural test, and Izod impact test, respectively.</p>
Full article ">Figure 9
<p>Schematic diagram of the strong interfacial bonding-enhanced rigid particle reinforcing mechanism of ADC particles in ADC-reinforced EMCs.</p>
Full article ">Figure 10
<p>Typical weight loss (<b>A</b>) and the derivative weight loss (<b>B</b>) against temperature of neat epoxy (control sample) and ADC-reinforced EMCs with ADC loading levels at 0.25 wt.%, 0.5 wt.%, 1 wt.%, and 2 wt.% from thermogravimetric analysis (TGA).</p>
Full article ">
31 pages, 8491 KiB  
Article
Characterisation and Application of Bio-Inspired Hybrid Composite Sensors for Detecting Barely Visible Damage under Out-of-Plane Loadings
by Ali Tabatabaeian, Reza Mohammadi, Philip Harrison and Mohammad Fotouhi
Sensors 2024, 24(16), 5170; https://doi.org/10.3390/s24165170 (registering DOI) - 10 Aug 2024
Viewed by 187
Abstract
Abstract: Traditional inspection methods often fall short in detecting defects or damage in fibre-reinforced polymer (FRP) composite structures, which can compromise their performance and safety over time. A prime example is barely visible impact damage (BVID) caused by out-of-plane loadings such as [...] Read more.
Abstract: Traditional inspection methods often fall short in detecting defects or damage in fibre-reinforced polymer (FRP) composite structures, which can compromise their performance and safety over time. A prime example is barely visible impact damage (BVID) caused by out-of-plane loadings such as indentation and low-velocity impact that can considerably reduce the residual strength. Therefore, developing advanced visual inspection techniques is essential for early detection of defects, enabling proactive maintenance and extending the lifespan of composite structures. This study explores the viability of using novel bio-inspired hybrid composite sensors for detecting BVID in laminated FRP composite structures. Drawing inspiration from the colour-changing mechanisms found in nature, hybrid composite sensors composed of thin-ply glass and carbon layers are designed and attached to the surface of laminated FRP composites exposed to transverse loading. A comprehensive experimental characterisation, including quasi-static indentation and low-velocity impact tests alongside non-destructive evaluations such as ultrasonic C-scan and visual inspection, is conducted to assess the sensors’ efficacy in detecting BVID. Moreover, a comparison between the two transverse loading types, static indentation and low-velocity impact, is presented. The results suggest that integrating sensors into composite structures has a minimal effect on mechanical properties such as structural stiffness and energy absorption, while substantially improving damage visibility. Additionally, the influence of fibre orientation of the sensing layer on sensor performance is evaluated, and correlations between internal and surface damage are demonstrated. Full article
(This article belongs to the Special Issue Damage Assessment and Structural Health Monitoring of Composites)
10 pages, 3599 KiB  
Article
An Analysis of the Biocompatibility, Cytotoxicity, and Bone Conductivity of Polycaprolactone: An In Vivo Study
by Wâneza Dias Borges Hirsch, Alexandre Weber, Janaine Ferri, Adriana Etges, Paulo Inforçatti Neto, Frederico David Alencar de Sena Pereira and Cláiton Heitz
Polymers 2024, 16(16), 2271; https://doi.org/10.3390/polym16162271 (registering DOI) - 10 Aug 2024
Viewed by 330
Abstract
Background: Tissue engineering represents a promising field in regenerative medicine, with bioresorbable polymers such as polycaprolactone (PCL) playing a crucial role as scaffolds. These scaffolds support the growth and repair of tissues by mimicking the extracellular matrix. Objective: This study aimed to assess [...] Read more.
Background: Tissue engineering represents a promising field in regenerative medicine, with bioresorbable polymers such as polycaprolactone (PCL) playing a crucial role as scaffolds. These scaffolds support the growth and repair of tissues by mimicking the extracellular matrix. Objective: This study aimed to assess the in vivo performance of three-dimensional PCL scaffolds by evaluating their effects on bone repair in rat calvaria and the tissue reaction in subcutaneous implant sites, as well as their impact on major organs such as the kidneys, lungs, and liver. Methods: Three-dimensional scaffolds made of PCL were implanted in the subcutaneous tissue of rats’ backs and calvaria. Histological analyses were conducted to observe the bone repair process in calvaria and the tissue response in subcutaneous implant sites. Additionally, the kidneys, lungs, and livers of the animals were examined for any adverse tissue alterations. Results: The histological analysis of the bone repair in calvaria revealed newly formed bone growing towards the center of the defects. In subcutaneous tissues, a thin fibrous capsule with collagenous fibers enveloping the implant was observed in all animals, indicating a positive tissue response. Importantly, no harmful alterations or signs of inflammation, hyperplasia, metaplasia, dysplasia, or hemorrhage were detected in the kidneys, lungs, and liver. Conclusions: The findings demonstrate that PCL scaffolds produced through additive manufacturing are biocompatible, non-cytotoxic, and bioresorbable, promoting osteoconduction without adverse effects on major organs. Hence, PCL is confirmed as a suitable biomaterial for further studies in tissue engineering and regenerative medicine. Full article
(This article belongs to the Special Issue Advanced Biodegradable Polymer Scaffolds for Tissue Engineering II)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Incision in rat’s calvarium. (<b>B</b>) Bone defects prepared with bone trephine. (<b>C</b>) Experimental bone defect filled with polycaprolactone disc and empty control defect.</p>
Full article ">Figure 2
<p>(<b>A</b>) Incisions at midline on rat’s back. (<b>B</b>) Insertion of a polycaprolactone disc into surgical cavity. (<b>C</b>) Suture of dorsal tissues.</p>
Full article ">Figure 3
<p>Histologic images of new formed bone in defects containing biomaterial at 7 days (<b>A</b>), 21 days (<b>B</b>), 60 days (<b>C</b>), 90 days (<b>D</b>), and 120 days, showing the formation of a bone bridge (<b>E</b>). Areas of new bone formation (arrows).</p>
Full article ">Figure 4
<p>Histologic images of animals’ organs. Kidney with mild glomerular hypercellularity (<b>A</b>), kidney with vascular congestion and foci of capillary aggregates (<b>B</b>), liver with vascular and sinusoidal congestion (<b>C</b>), liver with cells presenting with macrovesicular steatosis (arrow) (<b>D</b>), lung with peribronchial lymphoid aggregates (<b>E</b>), and lung with mild alveolar septal thickening and vascular congestion (<b>F</b>).</p>
Full article ">Figure 5
<p>Histologic images of tissues adjacent to the disc implanted on animals’ backs at 60 days. Formation of a thin fibrous capsule involving the implant (<b>A</b>), detail of the fibrous capsule, with organized collagen fibers involving the implant (<b>B</b>,<b>C</b>).</p>
Full article ">Figure 6
<p>Histologic images of animals’ organs. Kidney with mild glomerular hypercellularity and vascular congestion (<b>A</b>), liver with vascular and sinusoidal congestion and cell presenting with macrovesicular steatosis (arrow) (<b>B</b>), and lung with peribronchial lymphoid agglomerates, mild alveolar septal thickening, and vascular congestion (<b>C</b>).</p>
Full article ">
38 pages, 5633 KiB  
Review
Fundamental Aspects of Stretchable Mechanochromic Materials: Fabrication and Characterization
by Christina Tang
Materials 2024, 17(16), 3980; https://doi.org/10.3390/ma17163980 (registering DOI) - 10 Aug 2024
Viewed by 194
Abstract
Mechanochromic materials provide optical changes in response to mechanical stress and are of interest in a wide range of potential applications such as strain sensing, structural health monitoring, and encryption. Advanced manufacturing such as 3D printing enables the fabrication of complex patterns and [...] Read more.
Mechanochromic materials provide optical changes in response to mechanical stress and are of interest in a wide range of potential applications such as strain sensing, structural health monitoring, and encryption. Advanced manufacturing such as 3D printing enables the fabrication of complex patterns and geometries. In this work, classes of stretchable mechanochromic materials that provide visual color changes when tension is applied, namely, dyes, polymer dispersed liquid crystals, liquid crystal elastomers, cellulose nanocrystals, photonic nanostructures, hydrogels, and hybrid systems (combinations of other classes) are reviewed. For each class, synthesis and processing, as well as the mechanism of color change are discussed. To enable materials selection across the classes, the mechanochromic sensitivity of the different classes of materials are compared. Photonic systems demonstrate high mechanochromic sensitivity (Δnm/% strain), large dynamic color range, and rapid reversibility. Further, the mechanochromic behavior can be predicted using a simple mechanical model. Photonic systems with a wide range of mechanical properties (elastic modulus) have been achieved. The addition of dyes to photonic systems has broadened the dynamic range, i.e., the strain over which there is an optical change. For applications in which irreversible color change is desired, dye-based systems or liquid crystal elastomer systems can be formulated. While many promising applications have been demonstrated, manufacturing uniform color on a large scale remains a challenge. Standardized characterization methods are needed to translate materials to practical applications. The sustainability of mechanochromic materials is also an important consideration. Full article
(This article belongs to the Section Smart Materials)
Show Figures

Figure 1

Figure 1
<p>Overview of mechanochromic dyes incorporated into polyurethane elastomer systems. (<b>A</b>) Bis(benzoxazolyl)stibene (BBS), an aggrechromic dye, blended with polyurethane. Visible changes were observed under UV light due to disaggregation of the dye upon stretching; changes in hue (images numbered with roman numerals for reference) and brightness were quantified on a chromaticity diagram. Adapted with permission from [<a href="#B15-materials-17-03980" class="html-bibr">15</a>]. (<b>B</b>) Bis-alkene functionalized spiropyran was incorporated into PDMS beads and the beads were blended into polyurethane. The resulting films turned from yellow to blue when stretched. Adapted with permission from [<a href="#B17-materials-17-03980" class="html-bibr">17</a>]. (<b>C</b>) Rhodamine was blended with polyurethane (self-healable) and layered with PVA/TiO<sub>2</sub>. The mechanochromic response was repeatable after self-healing as quantified by the change in relative fluorescence intensity at 0% and 60% strain. Adapted with permission from [<a href="#B18-materials-17-03980" class="html-bibr">18</a>]. (<b>D</b>) Overview of synthesis of mechanochromic poly(ether-ester-urethane) elastomer containing spiropyran. The films turned from yellow to purple with strain. The average intensity of the blue channel B/(R + B + G) was used to monitor the onset of color activation. Adapted with permission from [<a href="#B19-materials-17-03980" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Overview of mechanochromic dyes incorporated into silicone elastomers (PDMS). (<b>A</b>) Spiropyran with alkene functionality was added as a crosslinker to commercially available silicone elastomer (e.g., Sylgard 184). Under strain, spiropyran (initially colorless) underwent a ring-opening transition to merocyanine (purple) as a visual indicator of strain. The color change was monitored by changes in the B/G channel intensity. Adapted with permission from [<a href="#B31-materials-17-03980" class="html-bibr">31</a>]. (<b>B</b>) Bis-alkene functionalized-spiropyran incorporated into PDMS. The original sample (colorless) turned blue when stretched, and purple when released to its initial shape. Arrows indicate ring opening. The color change was quantified by absorbance spectra. Adapted with permission from [<a href="#B32-materials-17-03980" class="html-bibr">32</a>]. (<b>C</b>) Mechanochromic PDMS network incorporated multiple dyes: naphthopyran (orange, slow kinetics) and spiropyran (purple, fast kinetics). Lines (red and blue dashed indicate models fit to the experimental data). Both dyes were activated when the sample was stretched resulting in a pink color. Spiropyran faded faster and after 2 min naphthopyran was activated and spiropyran faded, resulting in a transition from pink to orange. Adapted with permission from [<a href="#B33-materials-17-03980" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>Overview of mechanochromic polymer dispersed liquid crystals. (<b>A</b>) Fabrication process in which the liquid crystal (CLC) is mixed and emulsified with a polymer solution and cast into a film. Adapted with permission from [<a href="#B54-materials-17-03980" class="html-bibr">54</a>] (under open access CC-BY). (<b>B</b>) Using cholesteryl ester liquid formulations dispersed in polyurethane, films that changed color (e.g., green to blue upon initial stretch) were achieved. Adapted with permission from [<a href="#B55-materials-17-03980" class="html-bibr">55</a>]. (<b>C</b>) The mechano-optical response was characterized by UV reflectance as the sample was stretched and then held at 150% strain. A blue shift is observed upon initial stretching followed by a red shift. Adapted from with permission [<a href="#B55-materials-17-03980" class="html-bibr">55</a>]. (<b>D</b>) The proposed mechanism of color change is deformation of the liquid crystal droplets into an oblate shape upon mechanical strain. Adapted from [<a href="#B54-materials-17-03980" class="html-bibr">54</a>] (under open access CC-BY) and with permission from [<a href="#B56-materials-17-03980" class="html-bibr">56</a>].</p>
Full article ">Figure 4
<p>Overview of mechanochromic liquid crystal elastomers. (<b>A</b>) Opaque liquid crystal elastomer due to scattering of the liquid crystal domains. Upon stretching, the liquid crystal domains aligned resulting in increased transparency (adapted with permission from [<a href="#B60-materials-17-03980" class="html-bibr">60</a>]). Dotted portions are interpolated based on the plot. (<b>B</b>) Structurally colored liquid crystal elastomers obtained by injecting the components: liquid crystal monomer, chiral agent, crosslinker, plasticizer, and photoinitiator into a glass cell lined with PDMS via capillary force, applying shear force for orientation of the cholesteric phase, followed by photopolymerization. The resulting PDMS/liquid crystal elastomer film changes color when stretched (adapted with permission from [<a href="#B61-materials-17-03980" class="html-bibr">61</a>]). (<b>C</b>) Structurally colored liquid crystal elastomer films prepared by anisotropic deswelling in which the components are mixed in a solvent and cast on a glass substrate. A film that initially appears red turns blue upon stretching, confirmed with changes in the reflective spectra (adapted with permission from [<a href="#B62-materials-17-03980" class="html-bibr">62</a>]). (<b>D</b>) Liquid crystal elastomers prepared on PVA-coated substrates with photomask and multiple crosslinking steps to achieve colored patterns and transparent portions (adapted with permission from [<a href="#B63-materials-17-03980" class="html-bibr">63</a>]). (<b>E</b>) Using multiple crosslinking steps and a photomask, the colorless and transparent films could be achieved. Color appeared in the specified pattern upon stretching (adapted with permission from [<a href="#B64-materials-17-03980" class="html-bibr">64</a>]).</p>
Full article ">Figure 5
<p>(<b>A</b>) Overview of liquid crystal elastomer fibers prepared by anisotropic deswelling via extrusion onto a PVA-coated mandrel. (<b>B</b>) Images of mechanochromic response of resulting fibers under increasing strain observed under reflection mode polarized optical microscopy (200 µm scale bar). (<b>C</b>) Reflectance spectra of the selectively reflected light obtained through a right-handed circular polarizer. (<b>D</b>) Liquid crystal fiber (red) with increasing concentrations of Sudan black against white paper and black cloth under ambient light. (<b>E</b>) Resulting fibers woven (pain weave or hand sewn) into black elastic fabric under ambient light with 1 cm scale bars. Adapted with permission from [<a href="#B73-materials-17-03980" class="html-bibr">73</a>].</p>
Full article ">Figure 6
<p>Overview of mechanochromic materials using cellulose nanocrystals. (<b>A</b>) Self-assembly of photonic cellulose nanocrystals with glucose and polymer precursors followed by photopolymerization. SEM images of the films demonstrating the layer spacing on the order of visible light. Adapted with permission from [<a href="#B78-materials-17-03980" class="html-bibr">78</a>]. Representative mechanochromic response of cellulose nanocrystal/glucose/polyacrylate films. Adapted with permission from [<a href="#B75-materials-17-03980" class="html-bibr">75</a>]. (<b>B</b>) Shear alignment of cellulose nanocrystals dispersed in monomers followed by photopolymerization and swelling to achieving structurally colored mechanochromic films that changed when stretched (viewed under crossed polarizers with a 1 cm scale bar). Adapted with permission from [<a href="#B84-materials-17-03980" class="html-bibr">84</a>].</p>
Full article ">Figure 7
<p>Overview of photonic mechanochromic materials based on colloidal particle assemblies. (<b>A</b>) Deposition of monodisperse particles (silica or polystyrene) results in self-assembly of the particles into a colloidal array (ordered hexagonal arrangement). The colloidal array is infiltrated with monomers followed by polymerization, resulting in structurally colored, responsive composites (adapted with permission from [<a href="#B103-materials-17-03980" class="html-bibr">103</a>]). Representative SEM of the particle packing adapted from [<a href="#B104-materials-17-03980" class="html-bibr">104</a>] with permission. (<b>B</b>) Effect of particle volume fraction on resulting structural color for 150 nm (diameter) silica particles dispersed in PEGPEA (adapted with permission from [<a href="#B96-materials-17-03980" class="html-bibr">96</a>]). The effect of particle size on resulting structural color for silica particles dispersed in PEGPEA. The use of polydopamine particles (PDA) increased brightness by absorbing diffusely scattered light (adapted with permission from [<a href="#B95-materials-17-03980" class="html-bibr">95</a>]). (<b>C</b>) Dynamic mechanochromic response of photonic composite materials. The color changed due to a reduction in lattice spacing upon stretching. Adapted with permission from [<a href="#B103-materials-17-03980" class="html-bibr">103</a>]. (<b>D</b>) Colloidal assemblies patterned using custom designed templates demonstrating high mechanochromic sensitivity due to stress concentration in the patterned areas. Adapted with permission from [<a href="#B105-materials-17-03980" class="html-bibr">105</a>]. (<b>E</b>) Structurally colored, mechanochromic fibers produced from monodisperse silica using microfluidics. The fibers turned from red to green when stretched. Adapted with permission from [<a href="#B106-materials-17-03980" class="html-bibr">106</a>]. (<b>F</b>) Three-layer particles (polystyrene core and polymethacryate interlayer/shell) formulated for 3D printing. An initially red 3D printed dog bone turned from red to green upon increasing the strain from 0 to 30%. Adapted with permission from [<a href="#B98-materials-17-03980" class="html-bibr">98</a>].</p>
Full article ">Figure 8
<p>Overview of photonic metamaterials with mechanochromic properties. (<b>A</b>) Molybdenum disulfide, MoS<sub>2</sub>-based metamaterial from deposition of MoS<sub>2</sub> (strain-sensitive refractive index) on a nanograting. Changes in transmission were observed upon applying strain. Arrow indicates spectra with increasing strain. Adapted with permission from [<a href="#B114-materials-17-03980" class="html-bibr">114</a>]. (<b>B</b>) Photonic PDMS structure with periodic arrays of cylindrical holes prepared using a mold fabricated using nanolithography. The photonic material was layered on top of black PDMS for visual strain sensing. Adapted with permission from [<a href="#B88-materials-17-03980" class="html-bibr">88</a>]. (<b>C</b>) Patterning of photonic nanostructures on commercially available elastomeric photopolymer was achieved using a standard light project and reflecting surface. The resulting structurally colored image was bonded to black silicone; complex patterns such as flowers have been achieved. Materials were mechanochromic showing a blue shift with applied strain. Adapted from [<a href="#B116-materials-17-03980" class="html-bibr">116</a>] (CC-BY-SA-2.0).</p>
Full article ">Figure 9
<p>Overview of mechanochromic hydrogels. (<b>A</b>) Surfactant molecules self-assembled into two-dimensional bilayer lamellar structures and embedded in a cross-linked hydrogel matrix. When swollen in an appropriate solvent, reflective color was achieved. (<b>B</b>) Representative mechanochromic properties demonstrating visible color change from red to blue when stretched from 0 to 100% strain (confirmed with reflectance spectra) (1 cm scale bar). The blue shift was attributed to a decrease in lamellar spacing upon stretching. Arrows indicate structural color due to Bragg reflection. Adapted with permission from [<a href="#B118-materials-17-03980" class="html-bibr">118</a>].</p>
Full article ">Figure 10
<p>Overview of hybrid mechanochromic materials. (<b>A</b>) Porous PDMS-microparticle dye composites with hierarchical structure that turns from yellow to blue with strain due to spiropyran that shows an enhanced color transition region compared to dye in PDMS quantified using chromaticity. Adapted with permission from [<a href="#B120-materials-17-03980" class="html-bibr">120</a>]. (<b>B</b>) Dual mode mechanoresponsive material that integrated photonic crystal (silica particles in PEG base layer) and absorptive dye (PDMS-spiropyran). Color change with strain due to the dye and photonic crystal was observed in the reflectance spectra (black arrow dotted arrow indicates spectra taken at increasing strain, red arrows highlight changes in the spectra). Using hyperspectral imaging, complex strain and stress distributions around defects were visualized. Adapted with permission from [<a href="#B119-materials-17-03980" class="html-bibr">119</a>].</p>
Full article ">Figure 11
<p>Comparison of mechanochromic sensitivity (Δnm/%strain) as a function of elastic modulus (MPa) for various classes of mechanochromic materials Each class of material is grouped by the dotted lines to guide the eye [<a href="#B1-materials-17-03980" class="html-bibr">1</a>,<a href="#B6-materials-17-03980" class="html-bibr">6</a>,<a href="#B17-materials-17-03980" class="html-bibr">17</a>,<a href="#B32-materials-17-03980" class="html-bibr">32</a>,<a href="#B40-materials-17-03980" class="html-bibr">40</a>,<a href="#B48-materials-17-03980" class="html-bibr">48</a>,<a href="#B49-materials-17-03980" class="html-bibr">49</a>,<a href="#B55-materials-17-03980" class="html-bibr">55</a>,<a href="#B62-materials-17-03980" class="html-bibr">62</a>,<a href="#B64-materials-17-03980" class="html-bibr">64</a>,<a href="#B66-materials-17-03980" class="html-bibr">66</a>,<a href="#B67-materials-17-03980" class="html-bibr">67</a>,<a href="#B68-materials-17-03980" class="html-bibr">68</a>,<a href="#B73-materials-17-03980" class="html-bibr">73</a>,<a href="#B75-materials-17-03980" class="html-bibr">75</a>,<a href="#B78-materials-17-03980" class="html-bibr">78</a>,<a href="#B80-materials-17-03980" class="html-bibr">80</a>,<a href="#B81-materials-17-03980" class="html-bibr">81</a>,<a href="#B83-materials-17-03980" class="html-bibr">83</a>,<a href="#B93-materials-17-03980" class="html-bibr">93</a>,<a href="#B95-materials-17-03980" class="html-bibr">95</a>,<a href="#B96-materials-17-03980" class="html-bibr">96</a>,<a href="#B97-materials-17-03980" class="html-bibr">97</a>,<a href="#B100-materials-17-03980" class="html-bibr">100</a>,<a href="#B102-materials-17-03980" class="html-bibr">102</a>,<a href="#B103-materials-17-03980" class="html-bibr">103</a>,<a href="#B104-materials-17-03980" class="html-bibr">104</a>,<a href="#B105-materials-17-03980" class="html-bibr">105</a>,<a href="#B107-materials-17-03980" class="html-bibr">107</a>,<a href="#B108-materials-17-03980" class="html-bibr">108</a>,<a href="#B110-materials-17-03980" class="html-bibr">110</a>,<a href="#B111-materials-17-03980" class="html-bibr">111</a>,<a href="#B112-materials-17-03980" class="html-bibr">112</a>,<a href="#B113-materials-17-03980" class="html-bibr">113</a>,<a href="#B115-materials-17-03980" class="html-bibr">115</a>,<a href="#B118-materials-17-03980" class="html-bibr">118</a>,<a href="#B119-materials-17-03980" class="html-bibr">119</a>,<a href="#B124-materials-17-03980" class="html-bibr">124</a>,<a href="#B125-materials-17-03980" class="html-bibr">125</a>,<a href="#B126-materials-17-03980" class="html-bibr">126</a>,<a href="#B127-materials-17-03980" class="html-bibr">127</a>].</p>
Full article ">
12 pages, 4010 KiB  
Article
Improving Shale Stability through the Utilization of Graphene Nanopowder and Modified Polymer-Based Silica Nanocomposite in Water-Based Drilling Fluids
by Yerlan Kanatovich Ospanov, Gulzhan Abdullaevna Kudaikulova, Murat Smanovich Moldabekov and Moldir Zhumabaevna Zhaksylykova
Processes 2024, 12(8), 1676; https://doi.org/10.3390/pr12081676 (registering DOI) - 10 Aug 2024
Viewed by 198
Abstract
Shale formations present significant challenges to traditional drilling fluids due to fluid infiltration, cuttings dispersion, and shale swelling, which can destabilize the wellbore. While oil-based drilling fluids (OBM) effectively address these concerns about their environmental impact, their cost limits their widespread use. Recently, [...] Read more.
Shale formations present significant challenges to traditional drilling fluids due to fluid infiltration, cuttings dispersion, and shale swelling, which can destabilize the wellbore. While oil-based drilling fluids (OBM) effectively address these concerns about their environmental impact, their cost limits their widespread use. Recently, nanomaterials (NPs) have emerged as a promising approach in drilling fluid technology, offering an innovative solution to improve the efficiency of water-based drilling fluids (WBDFs) in shale operations. This study evaluates the potential of utilizing modified silica nanocomposite and graphene nanopowder to formulate a nanoparticle-enhanced water-based drilling fluid (NP-WBDF). The main objective is to investigate the impact of these nanoparticle additives on the flow characteristics, filtration efficiency, and inhibition properties of the NP-WBDF. In this research, a silica nanocomposite was successfully synthesized using emulsion polymerization and analyzed using FTIR, PSD, and TEM techniques. Results showed that the silica nanocomposite exhibited a unimodal particle size distribution ranging from 38 nm to 164 nm, with an average particle size of approximately 72 nm. Shale samples before and after interaction with the graphene nanopowder WBDF and the silica nanocomposite WBDF were analyzed using scanning electron microscopy (SEM). The NP-WBM underwent evaluation through API filtration tests (LTLP), high-temperature/high-pressure (HTHP) filtration tests, and rheological measurements conducted with a conventional viscometer. Experimental results showed that the silica nanocomposite and the graphene nanopowder effectively bridged and sealed shale pore throats, demonstrating superior inhibition performance compared to conventional WBDF. Post adsorption, the shale surface exhibited increased hydrophobicity, contributing to enhanced stability. Overall, the silica nanocomposite and the graphene nanopowder positively impacted rheological performance and provided favorable filtration control in water-based drilling fluids. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of modified polymer-based silica nanocomposite [<a href="#B16-processes-12-01676" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>SEM picture of SiO<sub>2</sub>-NPs (<b>a</b>) and graphene nanopowder (<b>b</b>).</p>
Full article ">Figure 3
<p>OFITE 800 rotational viscosimeter.</p>
Full article ">Figure 4
<p>OFITE HTHP filter press.</p>
Full article ">Figure 5
<p>OFITE dynamic linear swellmeter.</p>
Full article ">Figure 6
<p>FT-IR spectra of the silica nanocomposite.</p>
Full article ">Figure 7
<p>PSD analysis of the diluted silica nanocomposite.</p>
Full article ">Figure 8
<p>TEM image of the diluted silica nanocomposite.</p>
Full article ">Figure 9
<p>FESEM micrograph of WBDF: (<b>a</b>) the base WBDF; (<b>b</b>) the silica nanocomposite WBDF; (<b>c</b>) graphene nanopowder WBDF.</p>
Full article ">
19 pages, 5721 KiB  
Article
Novel Biobased Copolymers Based on Poly(butylene succinate) and Cutin: In Situ Synthesis and Structure Properties Investigations
by Evangelia D. Balla, Panagiotis A. Klonos, Apostolos Kyritsis, Monica Bertoldo, Nathanael Guigo and Dimitrios N. Bikiaris
Polymers 2024, 16(16), 2270; https://doi.org/10.3390/polym16162270 (registering DOI) - 10 Aug 2024
Viewed by 161
Abstract
The present work describes the synthesis of poly(butylene succinate) (PBSu)-cutin copolymers by the two-stage melt polycondensation method, esterification and polycondensation. Cutin was added in four different concentrations, 2.5, 5, 10, and 20 wt%, in respect to succinic acid. The obtained copolymers were studied [...] Read more.
The present work describes the synthesis of poly(butylene succinate) (PBSu)-cutin copolymers by the two-stage melt polycondensation method, esterification and polycondensation. Cutin was added in four different concentrations, 2.5, 5, 10, and 20 wt%, in respect to succinic acid. The obtained copolymers were studied using a variety of techniques such as Fourier transform infrared spectroscopy (FTIR), X-ray diffraction analysis (XRD), differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), polarized light microscopy (PLM), as well as diffuse reflectance spectroscopy (DRS). A series of results, in agreement between different techniques, revealed the formation of PBSu-cutin interactions, confirming indirectly the successful in situ synthetic route of copolymers. DSC and XRD combined with PLM results provided indications that the crystallization temperature increases with the addition of small amounts of cutin and gradually decreases with increasing concentration. The crystallization process was easier and faster at 2.5%, 5%, and 10% concentrations, whereas at 20%, it was comparable to neat PBSu. The presence of cutin, in general, leads to the facilitated crystallizability of PBSu (direct effect), whereas a moderate drop in the glass transition temperature is recorded, the latter being an indirect effect of cutin via crystallization. The thermal stability improved in the copolymers compared to neat PBSu. Water contact angle measurements confirmed that the addition of cutin decreased the hydrophilicity. The local and segmental relaxation mapping is demonstrated for PBSu/cutin here for the first time. Enzymatic hydrolysis and soil degradation tests showed that, overall, cutin accelerated the decomposition of the polymers. The copolymers may be proven useful in several applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesis of PBSu using succinic acid and 1.4-butanediol.</p>
Full article ">Figure 2
<p>Synthesis of PBSu/cutin polyesters using succinic acid and 1.4-butanediol and cutin.</p>
Full article ">Figure 3
<p>Gradual coloration of the obtained products due to cutin.</p>
Full article ">Figure 4
<p>(<b>Left</b>) ATR-FTIR spectroscopy, (<b>right</b>) details on the ester group vibration peak.</p>
Full article ">Figure 5
<p>XRD profiles of PBSu/cutin copolyesters.</p>
Full article ">Figure 6
<p>DSC (<b>a</b>) cooling and subsequent (<b>b</b>) heating traces for PBSu and the conditions described within the figures.</p>
Full article ">Figure 7
<p>Comparative DSC traces for all PBSu-based systems during (<b>a</b>) cooling at 20 °C/min and (<b>b</b>) subsequent cooling at 10 °C/min. The insets to (<b>b</b>) provide details of temperature regions of glass transition and melting. For comparison, included are the results.</p>
Full article ">Figure 8
<p>DRS results of ε′′(f,T) for (<b>a</b>) neat PBSu and (<b>b</b>) PBSu_5% cutin. Highlighted are selected isothermal curves, whereas marked are the recorded relaxation processes (peaks).</p>
Full article ">Figure 9
<p>Selected DRS results in the form of comparative ε′′(f) for all systems at (<b>a</b>) −80 °C regarding local mobility and (<b>b</b>) −20 °C regarding segmental mobility.</p>
Full article ">Figure 10
<p>TGA-dTG analysis of the prepared copolyesters.</p>
Full article ">Figure 11
<p>Contact angle measurements of the prepared materials.</p>
Full article ">Figure 12
<p>PLM photos of prepared PBSu/cutin polyesters.</p>
Full article ">Figure 13
<p>Weight loss during enzymatic hydrolysis (<b>left</b>), and soil degradation (<b>right</b>).</p>
Full article ">Figure 14
<p>SEM images of PBSu/cutin copolymers before (<b>a</b>–<b>c</b>) and after (<b>d</b>–<b>f</b>) 30 days of enzymatic hydrolysis, and before (<b>a</b>–<b>c</b>) and after (<b>g</b>–<b>i</b>) 90 days of soil degradation.</p>
Full article ">Figure 15
<p>DSC scans of PBSu neat, PBSu cutin 2.5, 5, 10, 20 wt%.</p>
Full article ">Figure 15 Cont.
<p>DSC scans of PBSu neat, PBSu cutin 2.5, 5, 10, 20 wt%.</p>
Full article ">
13 pages, 2670 KiB  
Review
Advances in Regenerative and Reconstructive Medicine in the Prevention and Treatment of Bone Infections
by Leticia Ramos Dantas, Gabriel Burato Ortis, Paula Hansen Suss and Felipe Francisco Tuon
Biology 2024, 13(8), 605; https://doi.org/10.3390/biology13080605 (registering DOI) - 10 Aug 2024
Viewed by 158
Abstract
Reconstructive and regenerative medicine are critical disciplines dedicated to restoring tissues and organs affected by injury, disease, or congenital anomalies. These fields rely on biomaterials like synthetic polymers, metals, ceramics, and biological tissues to create substitutes that integrate seamlessly with the body. Personalized [...] Read more.
Reconstructive and regenerative medicine are critical disciplines dedicated to restoring tissues and organs affected by injury, disease, or congenital anomalies. These fields rely on biomaterials like synthetic polymers, metals, ceramics, and biological tissues to create substitutes that integrate seamlessly with the body. Personalized implants and prosthetics, designed using advanced imaging and computer-assisted techniques, ensure optimal functionality and fit. Regenerative medicine focuses on stimulating natural healing mechanisms through cellular therapies and biomaterial scaffolds, enhancing tissue regeneration. In bone repair, addressing defects requires advanced solutions such as bone grafts, essential in medical and dental practices worldwide. Bovine bone scaffolds offer advantages over autogenous grafts, reducing surgical risks and costs. Incorporating antimicrobial properties into bone substitutes, particularly with metals like zinc, copper, and silver, shows promise in preventing infections associated with graft procedures. Silver nanoparticles exhibit robust antimicrobial efficacy, while zinc nanoparticles aid in infection prevention and support bone healing; 3D printing technology facilitates the production of customized implants and scaffolds, revolutionizing treatment approaches across medical disciplines. In this review, we discuss the primary biomaterials and their association with antimicrobial agents. Full article
Show Figures

Figure 1

Figure 1
<p>A diagram demonstrating multiple options for doping bone grafts or polymers for 3D printing using metal nanoparticles or antibiotics in bone reconstruction.</p>
Full article ">Figure 2
<p>Silver nanoparticles on bone surface used for orthopedic graft.</p>
Full article ">Figure 3
<p>Antibiotic-impregnated PLA models with <span class="html-italic">Staphylococcus aureus</span> test.</p>
Full article ">Figure 4
<p>Implants with PLA impregnated with antibiotics tested during surgery for hip replacement.</p>
Full article ">
Back to TopTop