[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (920)

Search Parameters:
Keywords = crosslink density

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 5081 KiB  
Article
Radially and Axially Oriented Ammonium Alginate Aerogels Modified with Clay/Tannic Acid and Crosslinked with Glutaraldehyde
by Lucía G. De la Cruz, Tobias Abt, Noel León and Miguel Sánchez-Soto
Gels 2024, 10(8), 526; https://doi.org/10.3390/gels10080526 (registering DOI) - 10 Aug 2024
Viewed by 311
Abstract
Lightweight materials that combine high mechanical strength, insulation, and fire resistance are of great interest to many industries. This work explores the properties of environmentally friendly alginate aerogel composites as potential sustainable alternatives to petroleum-based materials. This study analyzes the effects of two [...] Read more.
Lightweight materials that combine high mechanical strength, insulation, and fire resistance are of great interest to many industries. This work explores the properties of environmentally friendly alginate aerogel composites as potential sustainable alternatives to petroleum-based materials. This study analyzes the effects of two additives (tannic acid and montmorillonite clay), the orientation that results during casting, and the crosslinking of the biopolymer with glutaraldehyde on the properties of the aerogel composites. The prepared aerogels exhibited high porosities between 90% and 97% and densities in the range of 0.059–0.191 g/cm3. Crosslinking increased the density and resulted in excellent performance under loading conditions. In combination with axial orientation, Young’s modulus and yield strength reached values as high as 305 MPa·cm3/g and 7 MPa·cm3/g, respectively. Moreover, the alginate-based aerogels exhibited very low thermal conductivities, ranging from 0.038 W/m·K to 0.053 W/m·K. Compared to pristine alginate, the aerogel composites’ thermal degradation rate decreased substantially, enhancing thermal stability. Although glutaraldehyde promoted combustion, the non-crosslinked aerogel composites demonstrated high fire resistance. No flame was observed in these samples under cone calorimeter radiation, and a minuscule peak of heat release of 21 kW/m2 was emitted as a result of their highly efficient graphitization and fire suppression. The combination of properties of these bio-based aerogels demonstrates their potential as substituents for their fossil-based counterparts. Full article
Show Figures

Figure 1

Figure 1
<p>Aerogel preparation via the sol–gel method through radial and axial freeze casting (Created with Biorender.com Agrmt No. BG27599XKD).</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra of pristine ammonium alginate (AA) aerogel before and after its modification with TA and MMT and crosslinking with GTA; (<b>b</b>) XPS spectra of non-crosslinked A5 and crosslinked A5* and A5C5T2* aerogels.</p>
Full article ">Figure 3
<p>(<b>a</b>) Comparison of bulk and relative density and porosity of ammonium alginate aerogel composites: N<sub>2</sub> adsorption/desorption isotherms and BET specific surface of (<b>b</b>) A5C5-R and (<b>c</b>) A5C5-X aerogels and (<b>d</b>) pore volume and pore size distribution of A5C5-R and A5C5-X aerogels.</p>
Full article ">Figure 4
<p>Virtual µ-CT image of pristine alginate aerogel frozen in the (<b>a</b>) radial (A5-R) and (<b>b</b>) axial (A5-X) directions and (<b>c</b>) GTA-crosslinked alginate axial aerogel composite (A5C5T2-X*).</p>
Full article ">Figure 5
<p>SEM images revealing the evolution of aerogel structures as the solid content increased: (<b>a</b>) A5-R, (<b>b</b>) A5C5-R, and (<b>c</b>) A5C5T2-R; pore alignment in the axial orientation: (<b>d</b>) A5-X, (<b>e</b>)A5C5-X, and (<b>f</b>) A5C5T2-X, and (<b>h</b>) after GTA crosslinking. (<b>g</b>) EDS of the dispersion of MMT clay on the A5C5T2-X* aerogel.</p>
Full article ">Figure 6
<p>(<b>a</b>) Stress–strain compressive plots of alginate–clay–tannic acid aerogels; (<b>b</b>) specific modulus and specific yield stress of the aerogels studied.</p>
Full article ">Figure 7
<p>(<b>a</b>) Correlation between the thermal conductivities and bulk densities of the AA aerogel composites, (<b>b</b>) Scheme of different contributions in thermal conductivity and thermography of A5C5-R and A5C5T1-X aerogels on a hot plate surface at 100 °C, and (<b>c</b>) Thermal conductivity and effusivity of AA composite aerogels.</p>
Full article ">Figure 8
<p>(<b>a</b>) TGA weight loss; (<b>b</b>) derivative thermogravimetric curves of alginate composite aerogels.</p>
Full article ">Figure 9
<p>Main representative (<b>a</b>) HRR, (<b>b</b>) THR, and (<b>c</b>) ARHE curves from ammonium alginate composites; (<b>d</b>) photograph and SEM photomicrograph and EDS elemental mapping of the char of A5C5T2 after cone calorimetry; (<b>e</b>) Raman spectra of A5, A5C5, and A5C5T2 aerogel ashes.</p>
Full article ">Figure 10
<p>(<b>a</b>) Comparison of the compositions and properties of the aerogels evaluated in this study; (<b>b</b>) comparison of our ammonium alginate aerogels with other aerogel composites reported in the literature [<a href="#B49-gels-10-00526" class="html-bibr">49</a>,<a href="#B50-gels-10-00526" class="html-bibr">50</a>,<a href="#B51-gels-10-00526" class="html-bibr">51</a>,<a href="#B52-gels-10-00526" class="html-bibr">52</a>,<a href="#B53-gels-10-00526" class="html-bibr">53</a>,<a href="#B54-gels-10-00526" class="html-bibr">54</a>].</p>
Full article ">
15 pages, 4626 KiB  
Article
Shape-Memory Effect of 4D-Printed Gamma-Irradiated Low-Density Polyethylene
by Yunke Huang, Yongxiang Tao and Yan Wang
Crystals 2024, 14(8), 717; https://doi.org/10.3390/cryst14080717 (registering DOI) - 10 Aug 2024
Viewed by 188
Abstract
Four-dimensional-printed smart materials have a wide range of applications in areas such as biomedicine, aerospace, and soft robotics. Among 3D printing technologies, fused deposition molding (FDM) is economical, simple, and apply to thermoplastics. Cross-linked polyethylene (XLPE) forms a stable chemical cross-linking structure and [...] Read more.
Four-dimensional-printed smart materials have a wide range of applications in areas such as biomedicine, aerospace, and soft robotics. Among 3D printing technologies, fused deposition molding (FDM) is economical, simple, and apply to thermoplastics. Cross-linked polyethylene (XLPE) forms a stable chemical cross-linking structure and shows good shape-memory properties, but the sample is not soluble or fusible, which makes it hard to be applied in FDM printing. Therefore, in this work, a new idea of printing followed by irradiation was developed to prepare 4D-printed XLPE. First, low-density polyethylene (LDPE) was used to print the products using FDM technology and then cross-linked by gamma irradiation was used. The printing parameters were optimized, and the gel content, mechanical properties, and shape-memory behaviors were characterized. After gamma irradiation, the samples showed no new peak in FTIR spectra. And the samples exhibited good shape-memory capabilities. Increasing the irradiation dose increased the cross-linking degree and tensile strength and improved the shape-memory properties. However, it also decreased the elongation at break, and it did not affect the crystallization or melting behaviors of LDPE. With 120 kGy of irradiation, the shape recovery and fixity ratios (Rr and Rf) of the samples were 97.69% and 98.65%, respectively. After eight cycles, Rr and Rf remained at 96.30% and 97.76%, respectively, indicating excellent shape-memory performance. Full article
(This article belongs to the Section Organic Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>Preparation process of 4D printing shape-memory-cross-linked LDPE samples.</p>
Full article ">Figure 2
<p>Schematic diagram of printing orientations of FDM (<b>a</b>); deposition path of FDM filament (<b>b</b>).</p>
Full article ">Figure 3
<p>Digital photos of the LDPE dumbbell-shaped samples at different printing flows (<b>a</b>); SEM images of the cross-section of samples at 80% and 110% printing flows (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 4
<p>Digital photos of the LDPE dumbbell-shaped samples at different printing speeds (<b>a</b>); SEM images of the cross-section of samples at 35 mm/s and 15 mm/s printing speeds (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 5
<p>Digital photos of LDPE dumbbell-shaped samples at different printing temperatures (<b>a</b>); SEM images of the cross-section of samples at 140 °C and 200 °C printing temperatures (<b>b</b>,<b>c</b>).</p>
Full article ">Figure 6
<p>FTIR spectra of uncross-linked LDPE and 120 kGy cross-linked LDPE.</p>
Full article ">Figure 7
<p>The relationship between <span class="html-italic">R</span>(<span class="html-italic">s + s</span><sup>1/2</sup>) and <span class="html-italic">R</span><sup>1/2</sup> for the cross-linking printing samples.</p>
Full article ">Figure 8
<p>First cooling DSC curves and second heating DSC curves of the LDPE with different irradiation doses (<b>a</b>,<b>b</b>).</p>
Full article ">Figure 9
<p>Effect of irradiation doses on the mechanical properties of LDPE-printed samples.</p>
Full article ">Figure 10
<p>Shape-memory test results of samples with different irradiation doses.</p>
Full article ">Figure 11
<p>Effect of cycle times on shape-memory performance of LDPE-printed samples.</p>
Full article ">Figure 12
<p>Shape recovery process of the 4D-printed flower.</p>
Full article ">
12 pages, 3164 KiB  
Article
High-Temperature Water Electrolysis Properties of Membrane Electrode Assemblies with Nafion and Crosslinked Sulfonated Polyphenylsulfone Membranes by Using a Decal Method
by Je-Deok Kim
Membranes 2024, 14(8), 173; https://doi.org/10.3390/membranes14080173 - 8 Aug 2024
Viewed by 284
Abstract
To improve the stability of high-temperature water electrolysis, I prepared membrane electrode assemblies (MEAs) using a decal method and investigated their water electrolysis properties. Nafion 115 and crosslinked sulfonated polyphenylsulfone (CSPPSU) membranes were used. IrO2 was used as the oxygen evolution reaction [...] Read more.
To improve the stability of high-temperature water electrolysis, I prepared membrane electrode assemblies (MEAs) using a decal method and investigated their water electrolysis properties. Nafion 115 and crosslinked sulfonated polyphenylsulfone (CSPPSU) membranes were used. IrO2 was used as the oxygen evolution reaction (OER) catalyst, and Pt/C was used as the hydrogen evolution reaction (HER) catalyst. The conductivity of the CSPPSU membrane at 80 °C and 90% RH (relative humidity) is about four times lower than that of the Nafion 115 membrane. Single-cell water electrolysis was performed while measuring the current density and performing electrochemical impedance spectroscopy (EIS) at cell temperatures from 80 to 150 °C and the stability of the current density over time at 120 °C and 1.7 V. The current density of water electrolysis using Nafion 115 and CSPPSU membranes at 150 °C and 2 V was 1.2 A/cm2 for both. The current density of the water electrolysis using the CSPPSU membrane at 120 °C and 1.7 V was stable for 40 h. The decal method improved the contact between the CSPPSU membrane and the catalyst electrode, and a stable current density was obtained. Full article
(This article belongs to the Special Issue Membranes for Energy and the Environment)
Show Figures

Figure 1

Figure 1
<p>Photographs of a single cell, the anode side, and the cathode side; (<b>a</b>) SUS316L end plate, (<b>b</b>) Pt/Ti separator plate, and (<b>c</b>) carbon separator plate.</p>
Full article ">Figure 2
<p>Polarization curves: (<b>a</b>) Nafion 115 and (<b>b</b>) CSPPSU membranes at different operation temperatures.</p>
Full article ">Figure 3
<p>Electrochemical performance analysis of Nafion 115 membrane (<a href="#membranes-14-00173-f002" class="html-fig">Figure 2</a>a): (<b>a</b>) Polarization curves of the HFR-free cell; (<b>b</b>) HFR vs. current density; (<b>c</b>) HFR-free polarization data at low current densities, plotted on a logarithmic scale and (<b>d</b>) at current densities between 800 and 1200 mA/cm<sup>2</sup>.</p>
Full article ">Figure 4
<p>Electrochemical performance analysis of CSPPSU membrane (<a href="#membranes-14-00173-f002" class="html-fig">Figure 2</a>b): (<b>a</b>) Polarization curves of HFR-free cell; (<b>b</b>) HFR over current density; (<b>c</b>) HFR-free polarization data at low current densities, plotted on a logarithmic scale and (<b>d</b>) at current densities between 500 and 1200 mA/cm<sup>2</sup>.</p>
Full article ">Figure 5
<p>Nyquist plots measured at different operation temperatures of cells with (<b>a</b>) Nafion 115 and (<b>b</b>) CSPPSU membranes; (<b>c</b>) equivalent circuit used to fit the EIS data.</p>
Full article ">Figure 6
<p>Time dependence of the single cell with (<b>a</b>) Nafion 115 and (<b>b</b>) CSPPSU membranes at 120 °C and 1.7 V.</p>
Full article ">Figure 7
<p>Comparison of the time dependence of a single cell of an MEA with a Nafion 115 or a CSPPSU membrane made using a decal method and a porous IrO<sub>2</sub> electrode at 120 °C and 1.7 V [<a href="#B18-membranes-14-00173" class="html-bibr">18</a>,<a href="#B19-membranes-14-00173" class="html-bibr">19</a>].</p>
Full article ">
18 pages, 4956 KiB  
Article
An Exosome-Laden Hydrogel Wound Dressing That Can Be Point-of-Need Manufactured in Austere and Operational Environments
by E. Cate Wisdom, Andrew Lamont, Hannah Martinez, Michael Rockovich, Woojin Lee, Kristin H. Gilchrist, Vincent B. Ho and George J. Klarmann
Bioengineering 2024, 11(8), 804; https://doi.org/10.3390/bioengineering11080804 - 8 Aug 2024
Viewed by 397
Abstract
Skin wounds often form scar tissue during healing. Early intervention with tissue-engineered materials and cell therapies may promote scar-free healing. Exosomes and extracellular vesicles (EV) secreted by mesenchymal stromal cells (MSC) are believed to have high regenerative capacity. EV bioactivity is preserved after [...] Read more.
Skin wounds often form scar tissue during healing. Early intervention with tissue-engineered materials and cell therapies may promote scar-free healing. Exosomes and extracellular vesicles (EV) secreted by mesenchymal stromal cells (MSC) are believed to have high regenerative capacity. EV bioactivity is preserved after lyophilization and storage to enable use in remote and typically resource-constrained environments. We developed a bioprinted bandage containing reconstituted EVs that can be fabricated at the point-of-need. An alginate/carboxymethyl cellulose (CMC) biomaterial ink was prepared, and printability and mechanical properties were assessed with rheology and compression testing. Three-dimensional printed constructs were evaluated for Young’s modulus relative to infill density and crosslinking to yield material with stiffness suitable for use as a wound dressing. We purified EVs from human MSC-conditioned media and characterized them with nanoparticle tracking analysis and mass spectroscopy, which gave a peak size of 118 nm and identification of known EV proteins. Fluorescently labeled EVs were mixed to form bio-ink and bioprinted to characterize EV release. EV bandages were bioprinted on both a commercial laboratory bioprinter and a custom ruggedized 3D printer with bioprinting capabilities, and lyophilized EVs, biomaterial ink, and thermoplastic filament were deployed to an austere Arctic environment and bioprinted. This work demonstrates that EVs can be bioprinted with an alginate/CMC hydrogel and released over time when in contact with a skin-like substitute. The technology is suitable for operational medical applications, notably in resource-limited locations, including large-scale natural disasters, humanitarian crises, and combat zones. Full article
(This article belongs to the Section Biomedical Engineering and Biomaterials)
Show Figures

Figure 1

Figure 1
<p>Isolation and characterization of EVs from MSC-conditioned media. (<b>A</b>) Schematic of EV isolation using size exclusion chromatography followed by lyophilization and characterization using mass spectrometry, NTA, and TEM (<b>B</b>). NTA of EVs collected showed a particle size peak of 118 nm with a concentration of 6.67 × 10<sup>9</sup> +/− 1.69 × 10<sup>7</sup> EVs/mL. Fresh EVs (<b>C</b>) and EVs following lyophilization (<b>D</b>) were imaged with TEM. Red arrows are pointing to the fresh (<b>C</b>) and lyophilized EVs (<b>D</b>) imaged with TEM.</p>
Full article ">Figure 2
<p>Rheology of alginate/CMC dressing biomaterial ink. A 0.05–100% shear strain sweep was performed at 23 °C on a parallel plate rheometer with a 0.5 mm gap and 1 Hz frequency (<b>A</b>). Storage modulus, G′ (black squares) and loss modulus, G″ (black triangles) and plotted. The storage modulus value was independent of shear strain up to approximately 1%, and G′ is greater than G″ indicating the biomaterial ink is a viscoelastic solid. (<b>B</b>) Viscosity study where biomaterial ink was loaded on a parallel plate rheometer with a 0.5 mm gap, 1 Hz frequency, and a ramp logarithmic program for shear rate was used from 0.005 to 200 s<sup>−1</sup>. The shear rate is plotted versus shear stress at 10 °C (blue triangles), 23 °C (black triangles), and 37 °C (red triangles). Increasing temperature decreased the viscosity. The biomaterial ink is non-Newtonian and shear thinning at each temperature. (<b>C</b>) Biomaterial ink yield stress determination at 23 °C. Samples were loaded on a parallel plate rheometer with a 0.5 mm gap. Shear stress was varied from 1 to 300 Pa using a ramp linear program. The yield stress was calculated using rheometer software.</p>
Full article ">Figure 3
<p>Print fidelity of alginate/CMC biomaterial ink at two different infill percentages before and after crosslinking with CaCl<sub>2</sub>. The bio-ink was 3D printed into a 20 × 20 × 3 mm object using a commercial bioprinter (BioX, Cellink). Print parameters were 6 mm/sec and up to 100 kPa pressure using a 22-gauge conical tip (<b>top row</b>). Infill was either 20% or 10%. Following printing, the prints were incubated in CaCl<sub>2</sub> for 60 min to crosslink the alginate component (<b>bottom row</b>).</p>
Full article ">Figure 4
<p>Printed Alginate/CMC hydrogel dressings with varying crosslinking times. Hydrogel dressings were printed with 20% infill and crosslinked in CaCl<sub>2</sub> for 5, 15, or 60 min.</p>
Full article ">Figure 5
<p>Mechanical testing of the printed bandage dressing squares. (<b>A</b>) Linear portion of example stress–strain curves for samples with 20% infill crosslinked for 5 min, R = 0.99 (Black Diamonds), 15 min, R = 0.99 (Grey Squares) or 60 min, R = 0.99 (Black Circles). The data were fit to linear regression, and the slope of the curve fit is Young’s modulus. (<b>B</b>) Maximum force at 10% strain and Young’s modulus of printed alginate/CMC hydrogel dressings after 5, 15, and 60 min of crosslinking in CaCl<sub>2</sub>.</p>
Full article ">Figure 6
<p>Bioprinted alginate/CMC dressing with reconstituted, red fluorescently labeled EVs. The bioactive dressing was bioprinted with 20% infill and imaged before crosslinking (<b>A</b>) and after crosslinking (<b>B</b>). A confocal microscopy z-stack, tile scan at 20X magnification visualized as a 3D projection of the red-boxed region of the alginate/CMC/EV dressing (inset) (<b>C</b>). The labeled EVs are distributed throughout the hydrogel dressing material.</p>
Full article ">Figure 7
<p>Transfer of EVs from dressing bio-ink to collagen blocks. (<b>A</b>) A 3D-printed collagen block was used to simulate skin to test the transfer of EVs from the printed dressing to the skin. The alginate/CMC dressing was printed using 20% infill, crosslinked, and cut to 10 mm × 10 mm × 3 mm. It was placed on top of a similar-sized collagen block and incubated at 37 °C in a 6-well plate with 1 mL of PBS to keep the collagen hydrated. (<b>B</b>) Positive control: a solution of fluorescently labeled EVs pipetted on top of the collagen and left to absorb. The collagen block was removed at 24 h and imaged for the appearance of labeled EVs transferred from the dressings that were crosslinked for (<b>C</b>) 10 min or (<b>D</b>) 60 min. Microscopy images taken at 20X magnification.</p>
Full article ">Figure 8
<p>Three-dimensional printing and bioprinting of a wound dressing in a laboratory environment using a ruggedized 3D printer. The alginate/CMC dressing was bioprinted onto an FFF 3D printed PLA backing and crosslinked with CaCl<sub>2</sub> solution for 10 min.</p>
Full article ">Figure 9
<p>(<b>A</b>) CAD rendering and (<b>B</b>) bioprinted bandage with alginate/CMC EV-laden bio-ink dressing.</p>
Full article ">Figure 10
<p>Ruggedized 3D printer for point-of-need manufacturing of a bioactive wound dressing. The 3D printer contained three printheads. A fused filament fabrication (FFF) printhead was used to print the PLA thermoplastic backing. Two pneumatic printheads were used to print the alginate/EV bioactive bio-ink and a commercially available adhesive.</p>
Full article ">
16 pages, 8867 KiB  
Article
Structural, Thermal and Mechanical Assessment of Green Compounds with Natural Rubber
by Xavier Colom, Jordi Sans, Frederic de Bruijn, Fernando Carrillo and Javier Cañavate
Macromol 2024, 4(3), 566-581; https://doi.org/10.3390/macromol4030034 - 7 Aug 2024
Viewed by 196
Abstract
The inadequate disposal of tires poses a significant threat to human health and requires effective recycling solutions. The crosslinked structure of rubber, formed through sulfur bridges during vulcanization, presents a major challenge for recycling because it prevents the rubber scraps from being reshaped [...] Read more.
The inadequate disposal of tires poses a significant threat to human health and requires effective recycling solutions. The crosslinked structure of rubber, formed through sulfur bridges during vulcanization, presents a major challenge for recycling because it prevents the rubber scraps from being reshaped thermoplastically. Reclaiming or devulcanization aims to reverse this crosslinking, allowing waste rubber to be transformed into products that can be reprocessed and revulcanized, thereby saving costs and preserving resources. Microwave technology shows promise for devulcanization due to its ability to break sulfur crosslinks. In this study, we investigate the devulcanization of ground tire rubber (GTR) through a combined process applied to samples from both car and truck tires subjected to varying periods of microwave irradiation (0, 3, 5 and 10 min). The devulcanized GTR was then blended with natural rubber (NR) and underwent a new vulcanization process, simulating recycling for novel applications. The GTR was mixed with NR in proportions of 0, 10, 30 and 50 parts per hundred rubber (phr). This study also examines the differences between the GTR from car tires and GTR from truck tires. The results showed that the treatment effectively breaks the crosslinks in the GTR, creating double bonds (C=C) and improving the mechanical properties of the revulcanized samples. The crosslinking density and related properties of the samples increased with treatment time, reaching a maximum at 5 min of microwave treatment, followed by a decrease at 10 min. Additionally, the incorporation of GTR enhanced the thermal stability of the resulting materials. Full article
Show Figures

Figure 1

Figure 1
<p>Stress–strain curves for NR vulcanizates with different amounts (0,10, 30 and 50 phr) of non-devulcanized GTRcar (GTRc).</p>
Full article ">Figure 2
<p>Stress–strain curves for NR vulcanizates with different amounts (0,10, 30 and 50 phr) of non-devulcanized GTRtruck (GTRt).</p>
Full article ">Figure 3
<p>Tensile properties of NR/GTRcar compounds.</p>
Full article ">Figure 4
<p>Tensile properties of NR/GTRtruck compounds.</p>
Full article ">Figure 5
<p>Hardness average values of different samples.</p>
Full article ">Figure 6
<p>(<b>a</b>) Apparent crosslink density of GTRcar vulcanizate samples (µmol/cm<sup>3</sup>); (<b>b</b>) percentage (%) of mono-, di- and polysulfidic bonds correlated with apparent CLD for non-treated (NT) and TMmW5 compounds.</p>
Full article ">Figure 7
<p>(<b>a</b>) Apparent crosslink density of GTRtruck vulcanizate samples (µmol/cm<sup>3</sup>); (<b>b</b>) percentage (%) of mono-, di- and polysulfidic bonds correlated apparent CLD for non-treated (NT) and TMmW5 compounds.</p>
Full article ">Figure 8
<p>Spectral area of 100NR with 50 phr GTRcar devulcanized by microwaves for different periods of time: (<b>a</b>) NR, (<b>b</b>) 50GTRc(NT)/NR, (<b>c</b>) 50GTRc(TMmW3)/NR, (<b>d</b>) 50GTRc(TMmW5)/NR and (<b>e</b>) 50GTRc(TMmW10)/NR.</p>
Full article ">Figure 9
<p>Spectral area of 100NR with 50 phr GTRtruck devulcanized by microwaves for different periods of time: (<b>a</b>) NR, (<b>b</b>) 50GTRt(NT)/NR, (<b>c</b>) 50GTRt(TMmW3)/NR, (<b>d</b>) 50GTRt(TMmW5)/NR and (<b>e</b>) 50GTRt(TMmW10)/NR.</p>
Full article ">Figure 10
<p>SEM pictures of crystal morphology of ZnSt<sub>2</sub> on 0GTRc(NT)/NR surface: (<b>a</b>) (×500) and (<b>b</b>) (×8000).</p>
Full article ">Figure 11
<p>TGA and DTG curves of 100NR, 100NR10GTRc, 100NR30GTRc and 100NR50GTRc. All the samples have been treated for 10′ with microwaves.</p>
Full article ">Figure 12
<p>TGA and DTG curves of 100NR, 100NR10GTRt, 100NR30GTRt and 100NR50GTRt. All the samples have been treated for 10′ with microwaves.</p>
Full article ">Figure 13
<p>SEM microphotographs of the tensile fracture surface of the composites: (<b>A1</b>) natural rubber (×500); (<b>A2</b>) natural rubber (×50); (<b>B</b>) NR50GTRc(NT); (<b>C</b>) NR50GTRc10; (<b>D</b>) NR50GTRt(NT); (<b>E</b>) NR50GTRc10.</p>
Full article ">
13 pages, 4147 KiB  
Article
Synthesis of Flexible Polyamide Aerogels Cross-Linked with a Tri-Isocyanate
by Daniel A. Scheiman, Haiquan Guo, Katherine J. Oosterbaan, Linda McCorkle and Baochau N. Nguyen
Gels 2024, 10(8), 519; https://doi.org/10.3390/gels10080519 - 7 Aug 2024
Viewed by 247
Abstract
A new series of flexible polyamide (PA) aerogels was synthesized using terephthaloyl chloride (TPC), 2,2′-dimethylbenzidine (DMBZ) and cross-linked with an inexpensive, commercially available tri-isocyanate (Desmodur N3300A) at polymer concentrations of 6–8 wt.% total solids and repeating units, n, from 30 to 60. [...] Read more.
A new series of flexible polyamide (PA) aerogels was synthesized using terephthaloyl chloride (TPC), 2,2′-dimethylbenzidine (DMBZ) and cross-linked with an inexpensive, commercially available tri-isocyanate (Desmodur N3300A) at polymer concentrations of 6–8 wt.% total solids and repeating units, n, from 30 to 60. The cross-linked DMBZ-based polyamide aerogels obtained, after supercritically drying using liquid CO2, had shrinkages of 19–27% with densities ranging from 0.12 g/cm3 to 0.22 g/cm3, porosity and surface areas up to 91% and 309 m2/g, respectively, and modulus values ranging from 20.6 to 109 MPa. Evidence suggests that a higher flexibility could be achieved using DMBZ in the polyamide backbone with N3300A as a cross-linker, when compared to previously reported TPC-mPDA-BTC PA aerogels, N3300A-polyimide aerogels, and N3300-reinforced silica aerogels. Full article
(This article belongs to the Special Issue Aerogels—Preparation and Properties)
Show Figures

Figure 1

Figure 1
<p>Physical state of cross-linked PA aerogels made with (<b>a</b>) mPDA/TPC/N3300A (10 wt.%, <span class="html-italic">n</span> = 30), (<b>b</b>) ODA/TPC/BTC (10 wt.%, <span class="html-italic">n</span> = 30), and (<b>c</b>) DMBZ/TPC/N300A (8 wt.%, <span class="html-italic">n</span> = 30).</p>
Full article ">Figure 2
<p>A solid <sup>13</sup>C NMR of a polyamide aerogel formulated at 6 wt.% total polymer concentration with <span class="html-italic">n</span> of 30.</p>
Full article ">Figure 3
<p>FTIR spectrum of a polyamide aerogel formulated at 6 wt.% total polymer concentration with <span class="html-italic">n</span> of 30; (<b>a</b>) full scale from 4000 cm<sup>−1</sup> to 800 cm<sup>−1</sup> and (<b>b</b>) enlarged scale from 2000 cm<sup>−1</sup> to 800 cm<sup>−1</sup>.</p>
Full article ">Figure 4
<p>Empirical model for (<b>a</b>) shrinkage (%), (<b>b</b>) density, and (<b>c</b>) porosity (%) vs. polymer concentration and total <span class="html-italic">n</span>.</p>
Full article ">Figure 5
<p>SEM images of the N3300A cross-linked PA aerogels at 6 wt.% with (<b>a</b>) <span class="html-italic">n</span> = 30 (ρ = 12.1 mg/cm<sup>3</sup>), (<b>b</b>) <span class="html-italic">n</span> = 45 (ρ = 12.9 mg/cm<sup>3</sup>), (<b>c</b>) <span class="html-italic">n</span> = 60 (ρ = 13.1 mg/cm<sup>3</sup>), and at 8 wt.% with (<b>d</b>) <span class="html-italic">n</span> = 30 (ρ = 17.1 mg/cm<sup>3</sup>) (<b>e</b>) <span class="html-italic">n</span> = 45 (ρ = 16.5 mg/cm<sup>3</sup>) and (<b>f</b>) <span class="html-italic">n</span> = 60 (ρ = 22.1 mg/cm<sup>3</sup>).</p>
Full article ">Figure 6
<p>Pore volume vs. pore diameter of the cross-linked PA aerogels at different n values and polymer concentrations.</p>
Full article ">Figure 7
<p>Adsorption and desorption isotherms of the cross-linked PA aerogels at different n values and polymer concentrations.</p>
Full article ">Figure 8
<p>BET surface area of the cross-linked PA aerogels.</p>
Full article ">Figure 9
<p>(<b>a</b>) Typical stress–strain curves from compression test of N3300A-cross-linked PA aerogel; (<b>b</b>) log–log plot of density vs. Young’s modulus of N3300A-DMBZ-TPC (closed symbol), and BTC-mPDA-TPC (open symbol) [<a href="#B15-gels-10-00519" class="html-bibr">15</a>] polyamide aerogels, N3300A-DMBZ-BPDA polyimide aerogels [<a href="#B19-gels-10-00519" class="html-bibr">19</a>], and N3300A-reinforced silica aerogels [<a href="#B30-gels-10-00519" class="html-bibr">30</a>].</p>
Full article ">Figure 10
<p>TGA curves of N3300A cross-linked PA aerogels at <span class="html-italic">n</span> = 30 and <span class="html-italic">n</span> = 60.</p>
Full article ">Figure 11
<p>Empirical model of final onset decomposition temperature of N-3300A cross-linked polyamide aerogel.</p>
Full article ">Scheme 1
<p>Chemical reaction of a N3300A cross-linked polyamide aerogels.</p>
Full article ">Scheme 2
<p>Schematically synthetic steps in fabricating N3300A cross-linked polyamide (or x-linked PA) aerogels.</p>
Full article ">
15 pages, 6254 KiB  
Article
Effects of the Amylose/Amylopectin Ratio of Starch on Borax-Crosslinked Hydrogels
by Kai Lu, Rudy Folkersma, Vincent S. D. Voet and Katja Loos
Polymers 2024, 16(16), 2237; https://doi.org/10.3390/polym16162237 - 6 Aug 2024
Viewed by 391
Abstract
Herein, we simultaneously prepared borax-crosslinked starch-based hydrogels with enhanced mechanical properties and self-healing ability via a simple one-pot method. The focus of this work is to study the effects of the amylose/amylopectin ratio of starch on the grafting reactions and the performance of [...] Read more.
Herein, we simultaneously prepared borax-crosslinked starch-based hydrogels with enhanced mechanical properties and self-healing ability via a simple one-pot method. The focus of this work is to study the effects of the amylose/amylopectin ratio of starch on the grafting reactions and the performance of the resulting borax-crosslinked hydrogels. An increase in the amylose/ amylopectin ratio increased the gel fraction and grafting ratio but decreased the swelling ratio and pore diameter. Compared with hydrogels prepared from low-amylose starches, hydrogels prepared from high-amylose starches showed pronouncedly increased network strength, and the maximum storage modulus increased by 8.54 times because unbranched amylose offered more hydroxyl groups to form dynamic borate ester bonds with borate ions and intermolecular hydrogen bonds, leading to an enhanced crosslink density. In addition, all the hydrogels exhibited a uniformly interconnected network structure. Furthermore, owing to the dynamic borate ester bonds and hydrogen bonds, the hydrogel exhibited excellent recovery behavior under continuous step strain, and it also showed thermal responsiveness. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) FTIR spectra of starch and crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G); (<b>b</b>) XRD patterns of native starches and crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G); and (<b>c</b>) <sup>1</sup>H-NMR spectra of starch, starch-<span class="html-italic">g</span>-PAM, and crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G).</p>
Full article ">Figure 2
<p>SEM images and pore size distributions of crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G).</p>
Full article ">Figure 3
<p>(<b>a</b>) Gel fractions of crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G); (<b>b</b>) swelling ratios of crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G). Values with different letters are significantly different (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>TGA and DTG of native starches and crosslinked hydrogels (Waxy-G, Maize-G, G50-G, and G80-G).</p>
Full article ">Figure 5
<p>(<b>a</b>) Dynamic strain sweep curve at ω = 10 rad/s; (<b>b</b>) dynamic frequency sweep curve at γ = 1%.</p>
Full article ">Figure 6
<p>(<b>a</b>) Self-healing pictures of G80-G hydrogels; (<b>b</b>) <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″ versus time for original and self-healing G80-G and G80-CG hydrogels after being cut; and (<b>c</b>) continuous step strain measurements of the G80-G hydrogel at strains of 1% and 100%.</p>
Full article ">Figure 7
<p>Temperature dependence of the <span class="html-italic">G</span>′ and <span class="html-italic">G</span>″ for the G80-G hydrogel during a heating–cooling–heating cycle at ω = 10 rad/s and γ = 1%.</p>
Full article ">Scheme 1
<p>Method and mechanism of synthesis of borax-crosslinked hydrogels.</p>
Full article ">
25 pages, 4744 KiB  
Article
Tribomechanical Properties of PVA/Nomex® Composite Hydrogels for Articular Cartilage Repair
by Francisco Santos, Carolina Marto-Costa, Ana Catarina Branco, Andreia Sofia Oliveira, Rui Galhano dos Santos, Madalena Salema-Oom, Roberto Leonardo Diaz, Sophie Williams, Rogério Colaço, Célio Figueiredo-Pina and Ana Paula Serro
Gels 2024, 10(8), 514; https://doi.org/10.3390/gels10080514 - 3 Aug 2024
Viewed by 501
Abstract
Due to the increasing prevalence of articular cartilage diseases and limitations faced by current therapeutic methodologies, there is an unmet need for new materials to replace damaged cartilage. In this work, poly(vinyl alcohol) (PVA) hydrogels were reinforced with different amounts of Nomex® [...] Read more.
Due to the increasing prevalence of articular cartilage diseases and limitations faced by current therapeutic methodologies, there is an unmet need for new materials to replace damaged cartilage. In this work, poly(vinyl alcohol) (PVA) hydrogels were reinforced with different amounts of Nomex® (known for its high mechanical toughness, flexibility, and resilience) and sterilized by gamma irradiation. Samples were studied concerning morphology, chemical structure, thermal behavior, water content, wettability, mechanical properties, and rheological and tribological behavior. Overall, it was found that the incorporation of aramid nanostructures improved the hydrogel’s mechanical performance, likely due to the reinforcement’s intrinsic strength and hydrogen bonding to PVA chains. Additionally, the sterilization of the materials also led to superior mechanical properties, possibly related to the increased crosslinking density through the hydrogen bonding caused by the irradiation. The water content, wettability, and tribological performance of PVA hydrogels were not compromised by either the reinforcement or the sterilization process. The best-performing composite, containing 1.5% wt. of Nomex®, did not induce cytotoxicity in human chondrocytes. Plugs of this hydrogel were inserted in porcine femoral heads and tested in an anatomical hip simulator. No significant changes were observed in the hydrogel or cartilage, demonstrating the material’s potential to be used in cartilage replacement. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images of the surface and cross-section of hydrogel samples, acquired with 1000× magnification.</p>
Full article ">Figure 2
<p>Equilibrium water content (<b>A</b>) and water contact angles (<b>B</b>) for PVA and PMIA-reinforced hydrogels. Data followed a normal distribution: Welch’s ANOVA and Dunnett’s C test or Student’s <span class="html-italic">t</span> test, <span class="html-italic">p</span> ≤ 0.05 (*).</p>
Full article ">Figure 3
<p>Typical compressive stress–strain curves for PVA- and PMIA-reinforced hydrogels before (<b>A</b>) and after (<b>C</b>) sterilization and corresponding values of compressive tangent modulus between 5% and 35% (<b>B</b> and <b>D</b>, respectively).</p>
Full article ">Figure 4
<p>Typical tensile stress–strain curves of the PVA- and PMIA-reinforced hydrogels before and after sterilization.</p>
Full article ">Figure 5
<p>Tensile properties of PVA- and nanofiber-reinforced hydrogels: (<b>A</b>) tensile modulus; (<b>B</b>) UTS; (<b>C</b>) toughness; (<b>D</b>) elongation at break. When comparing non-irradiated vs. irradiated samples, data followed a normal distribution, except for tensile moduli and UTS: Student’s <span class="html-italic">t</span> test or Mann–Whitney U test. When comparing samples with different amounts of PMIA, data followed a normal distribution, except for tensile moduli: Welch’s ANOVA and Dunnett’s C test or Kruskal–Wallis tests adjusted with Bonferroni correction. <span class="html-italic">p</span> ≤ 0.05 (*), <span class="html-italic">p</span> ≤ 0.01 (**), <span class="html-italic">p</span> ≤ 0.001 (***).</p>
Full article ">Figure 6
<p>Storage (G′) and loss (G″) moduli of the PVA- and PMIA-reinforced hydrogels before (<b>A</b>) and after (<b>B</b>) sterilization as a function of the frequency (0.1–10 Hz).</p>
Full article ">Figure 7
<p>Friction coefficients of PVA- and PMIA-reinforced hydrogels before and after sterilization were measured against stainless steel 316L in PBS solution. Data followed a normal distribution: Welch’s ANOVA and Dunnett’s C test or Student’s <span class="html-italic">t</span> test. <span class="html-italic">p</span> ≤ 0.05 (*), <span class="html-italic">p</span> ≤ 0.01 (**).</p>
Full article ">Figure 8
<p>Friction coefficients of PVA10_G and PVA10/PMIA1.5_G were measured against porcine cartilage using PBS solution and SF fluid as lubricants. Data followed a normal distribution: Student’s <span class="html-italic">t</span> test, <span class="html-italic">p</span> ≤ 0.05 (*), <span class="html-italic">p</span> ≤ 0.01 (**), <span class="html-italic">p</span> ≤ 0.001 (***).</p>
Full article ">Figure 9
<p>Acetabular cup (<b>A</b>,<b>B</b>) and femoral head with PVA10/PMIA1.5_G hydrogel plug (<b>C</b>,<b>D</b>) before (<b>A</b>,<b>C</b>) and after (<b>B</b>,<b>D</b>) the tests in a hip movement simulator. SEM images of the hydrogel before (<b>E</b>) and after (<b>F</b>) the simulation test.</p>
Full article ">Figure 10
<p>Human chondrocyte cell viability after 24 h exposure to extracts of hydrogel: MTT assay (<b>A</b>), cell morphology of PVA10/PMIA1.5_G (<b>B</b>), the negative control (<b>C</b>), and positive control (<b>D</b>). For the negative control, the cells were cultured in DMEM/F12, while for the positive control, they were cultured in DMEM/F12 with 10% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) DMSO. Data followed a normal distribution: Welch’s ANOVA and Dunnett’s C test, <span class="html-italic">p</span> ≤ 0.05 (*).</p>
Full article ">Figure 11
<p>Chemical structure of poly(meta-phenylene isophthalamide (PMIA) and poly(vinyl alcohol) (PVA).</p>
Full article ">
15 pages, 6006 KiB  
Article
Recycling of Commercially Available Biobased Thermoset Polyurethane Using Covalent Adaptable Network Mechanisms
by Edoardo Miravalle, Gabriele Viada, Matteo Bonomo, Claudia Barolo, Pierangiola Bracco and Marco Zanetti
Polymers 2024, 16(15), 2217; https://doi.org/10.3390/polym16152217 - 3 Aug 2024
Viewed by 374
Abstract
Until recently, recycling thermoset polyurethanes (PUs) was limited to degrading methods. The development of covalent adaptable networks (CANs), to which PUs can be assigned, has opened novel possibilities for actual recycling. Most efforts in this area have been directed toward inventing new materials [...] Read more.
Until recently, recycling thermoset polyurethanes (PUs) was limited to degrading methods. The development of covalent adaptable networks (CANs), to which PUs can be assigned, has opened novel possibilities for actual recycling. Most efforts in this area have been directed toward inventing new materials that can benefit from CAN theory; presently, little or nothing has been applied to industrially producible materials. In this study, both an industrially available polyol (Sovermol780®) and isocyanate (Tolonate X FLO 100®) with percentages of bioderived components were employed, resulting in a potentially scalable and industrially producible material. The resultant network could be reworked up to three times, maintaining the crosslinked structure without significantly changing the thermal properties. Improvements in mechanical parameters were observed when comparing the pristine material to the material exposed to three rework processes, with gains of roughly 50% in elongation at break and 20% in tensile strength despite a 25% decrease in Young’s modulus and crosslink density. Thus, it was demonstrated that theory may be profitably applied even to materials that are not designed including additional bonds but instead rely just on the dynamic urethane bond that is naturally present in the network. Full article
(This article belongs to the Special Issue Polymers and Biopolymers for Sustainable Life and Applications)
Show Figures

Figure 1

Figure 1
<p>PU network through the reprocessing steps.</p>
Full article ">Figure 2
<p>FTIR-ATR spectra of polyol and isocyanate precursor used and the pristine PU network.</p>
Full article ">Figure 3
<p>Structure proposed for isocyanate precursor, Tolonate X FLO 100 [<a href="#B42-polymers-16-02217" class="html-bibr">42</a>].</p>
Full article ">Figure 4
<p>TGA and DTG profiles of precursors and the PU network.</p>
Full article ">Figure 5
<p>(<b>a</b>) Storage modulus of the pristine network in the range of 40–190 °C; (<b>b</b>) stress relaxation curves of the pristine network at 160 °C, 170 °C, and 180 °C; (<b>c</b>) Arrhenius analysis of the pristine network.</p>
Full article ">Figure 6
<p>Comparison between the pristine network (first image) and the first, second, and third reprocessing (following images).</p>
Full article ">Figure 7
<p>FTIR-ATR spectra of the pristine material and reprocessed samples.</p>
Full article ">Figure 8
<p>(<b>a</b>) FTIR-ATR spectra of the pristine network and the same network heated to reprocessing temperature; (<b>b</b>) urethane exchange mechanisms proposed.</p>
Full article ">Figure 9
<p>TGA and DTG profiles of the pristine network and reprocessed samples.</p>
Full article ">Figure 10
<p>Swelling degree and weight loss of the pristine network and reprocessed samples.</p>
Full article ">Figure 11
<p>Storage modulus of the pristine network and reprocessed samples in the range of 40–190 °C.</p>
Full article ">Figure 12
<p>(<b>a</b>) Representative stress–strain curves for each sample, (<b>b</b>) values of stress and elongation at break for each sample.</p>
Full article ">
17 pages, 2461 KiB  
Article
New Approaches for Basophil Activation Tests Employing Dendrimeric Antigen–Silica Nanoparticle Composites
by Silvia Calvo-Serrano, Esther Matamoros, Jose Antonio Céspedes, Rubén Fernández-Santamaría, Violeta Gil-Ocaña, Ezequiel Perez-Inestrosa, Cecilia Frecha, Maria I. Montañez, Yolanda Vida, Cristobalina Mayorga and Maria J. Torres
Pharmaceutics 2024, 16(8), 1039; https://doi.org/10.3390/pharmaceutics16081039 - 3 Aug 2024
Viewed by 410
Abstract
In vitro cell activation through specific IgE bound to high-affinity receptors on the basophil surface is a widely used strategy for the evaluation of IgE-mediated immediate hypersensitivity reactions to betalactams. Cellular activation requires drug conjugation to a protein to form a large enough [...] Read more.
In vitro cell activation through specific IgE bound to high-affinity receptors on the basophil surface is a widely used strategy for the evaluation of IgE-mediated immediate hypersensitivity reactions to betalactams. Cellular activation requires drug conjugation to a protein to form a large enough structure displaying a certain distance between haptens to allow the cross-linking of two IgE antibodies bound to the basophil’s surface, triggering their degranulation. However, no information about the size and composition of these conjugates is available. Routine in vitro diagnosis using the basophil activation test uses free amoxicillin, which is assumed to conjugate to a carrier present in blood. To standardize the methodology, we propose the use of well-controlled and defined nanomaterials functionalized with amoxicilloyl. Silica nanoparticles decorated with PAMAM–dendrimer–amoxicilloyl conjugates (NpDeAXO) of different sizes and amoxicilloyl densities (50–300 µmol amoxicilloyl/gram nanoparticle) have been prepared and chemically characterized. Two methods of synthesis were performed to ensure reproducibility and stability. Their functional effect on basophils was measured using an in-house basophil activation test (BAT) that determines CD63+ or CD203chigh activation markers. It was observed that NpDeAXO nanocomposites are not only able to specifically activate basophils but also do so in a more effective way than free amoxicillin, pointing to a translational potential diagnosis. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>General procedure for the chemical modification of ϕ<b>dNp</b> dispersions and <b>50Nps</b> with different DeAXO densities in their surface.</p>
Full article ">Figure 2
<p>NMR spectra of (<b>a</b>) AX in basic D<sub>2</sub>O and (<b>b</b>) <b>50NpDeAXO</b> in D<sub>2</sub>O suspensions.</p>
Full article ">Figure 3
<p>Basophil activation test (BAT) dose–response curves of <b>NpDeAXO</b> of different sizes: 20 nm (<b>A</b>,<b>B</b>), 30 nm (<b>C</b>,<b>D</b>), and 50 nm (<b>E</b>,<b>F</b>) and (<b>G</b>,<b>H</b>). <span class="html-italic">Np dispersions</span> or <span class="html-italic">solid-state Nps</span> labels at the top of the figure only indicate the synthetic methodology used for Np preparation. Black lines represent healthy controls (HCs), and blue and green lines represent allergic patients (APs). Size sample included HCs (N = 10) and APs (N = 10) for Nps synthesized as dispersions (<b>A</b>–<b>F</b>), while HCs (N = 45) and APs (N = 54) were included in the study for the Nps synthesized as a solid state. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>BAT dose–response curves of Nps with different DeAXO surface densities: 300 µmol AXO/gNp (<b>A</b>,<b>B</b>); 130 µmolAXO/gNp (<b>C</b>,<b>D</b>); 100 µmol AXO/gNp (<b>E</b>,<b>F</b>); 80 µmol AXO/gNp (<b>G</b>,<b>H</b>); 50 µmol AXO/gNp (<b>I</b>,<b>J</b>). Black lines represent healthy controls (HCs), and blue and green lines represent allergic patients (APs). Size sample included HCs (N = 6) and APs (N = 4) (<b>A</b>–<b>J</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>BAT dose–response curves using <b>50NpDeAXO</b> and free AX at three different concentrations. APs (N = 54) are depicted by blue or green lines and HCs (N = 45) are depicted by black lines. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Percentage of positive cases in allergic patients (N = 54) in BAT using CD63 (<b>A</b>) and CD203c<sup>high</sup> (<b>B</b>) as basophil activation markers. Positive cases were obtained after using the cut-offs described in <a href="#app1-pharmaceutics-16-01039" class="html-app">Table S4</a> for AX and <b>50NpDeAXO</b> at the different concentrations with both CD63 and CD203c.</p>
Full article ">
11 pages, 4437 KiB  
Article
Effects of Aging on New Bone Regeneration in a Mandibular Bone Defect in a Rat Model
by Jung Ho Park, Jong Hoon Park, Hwa Young Yu and Hyun Seok
Biomimetics 2024, 9(8), 466; https://doi.org/10.3390/biomimetics9080466 - 1 Aug 2024
Viewed by 365
Abstract
The effects of aging on the healing capacity of maxillofacial bone defects have not been studied. The aim of this study was to evaluate the effects of aging on the regeneration of round bony defects in the mandible. We created a round-shaped bony [...] Read more.
The effects of aging on the healing capacity of maxillofacial bone defects have not been studied. The aim of this study was to evaluate the effects of aging on the regeneration of round bony defects in the mandible. We created a round-shaped bony defect in the mandibular angle area in rats of different ages (2-[2 M], 10-[10 M], and 20-month-old [20 M]) and evaluated new bone regeneration in these groups. Changes in bone turnover markers such as alkaline phosphatase, procollagen type I N-terminal propeptide (PINP), cross-linked C-telopeptide of type I collagen, and tartrate-resistant acid phosphatase 5B (TRAP5b) were investigated. The bone volume/total volume and bone mineral density of the 20 M group were significantly higher than those of the 2 M group (p = 0.029, 0.019). A low level of the bone formation marker PINP was observed in the 20 M group, and a high level of the bone resorption marker TRAP5b was observed in the 10 M and 20 M groups. Micro-computed tomography (µ-CT) results showed that older rats had significantly higher bone formation than younger rats, with lower serum levels of PINP and higher levels of TRAP5b. The local environment of the old rat bone defects, surrounded by thickened bone, may have affected the results of our study. In conclusion, old rats showed greater new bone regeneration and healing capacity for round mandibular bone defects. This result was related to the fact that the bone defects in the 20 M rat group provided more favorable conditions for new bone regeneration. Full article
Show Figures

Figure 1

Figure 1
<p>Micro-computed tomography analysis. (<b>A</b>) Bone volume/total volume (BV/TV), (<b>B</b>) bone mineral density (BMD), (<b>C</b>) trabecular thickness (TbTh), (<b>D</b>) trabecular space (TbSp), and (<b>E</b>) total volume (TV) of 2 M, 10 M, and 20 M groups. There was a significant difference in the BV, BMD, and TV among the three groups (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>The sagittal images of the micro-computed tomography of the mandible bone defect of (<b>A</b>) 2 M, (<b>B</b>) 10 M, and (<b>C</b>) 20 M groups. The new bone regeneration was observed from the edge of the bone defect.</p>
Full article ">Figure 3
<p>Three-dimensional (3D) reconstruction image of micro-computed tomography (μ-CT) of the mandible bone of (<b>A</b>) 2 M, (<b>B</b>) 10 M, and (<b>C</b>) 20 M groups. The new bone regeneration was more prominent in the 20 M group.</p>
Full article ">Figure 4
<p>Serum levels of (<b>A</b>) ALP, (<b>B</b>) PINP, (<b>C</b>) CTXI, and (<b>D</b>) TRAP5b. PINP levels in the 2 M group were significantly higher than those in the 10 M and 20 M groups at T0 (<span class="html-italic">p</span> = 0.026 and 0.002, respectively) and the 20 M group at T2 (<span class="html-italic">p</span> = 0.006). PINP levels in the 10 M group were significantly higher than those in the 20 M group (<span class="html-italic">p</span> = 0.020) at T2. The serum TRAP5b levels in the 10 M group were significantly higher than those in the 2 M group at T0, T1, and T2 (<span class="html-italic">p</span> = 0.039, 0.016, and 0.024, respectively) (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Histological images (hematoxylin and eosin staining) of each group. (<b>A</b>) 2 M, (<b>B</b>) 10 M, and (<b>C</b>) 20 M groups. A part of the new bone was formed at the end edge of the bone defect in the 2 M group. A remarkable bone regeneration was formed and covered the bone defect in the 10 M group. The thick and mature bone regeneration was observed in the 20 M group (original magnification 4×, scale bar = 1 mm).</p>
Full article ">Figure 6
<p>Immunohistochemical staining of ALP (<b>A</b>–<b>C</b>) and TRAP (<b>D</b>–<b>F</b>) in each group. (<b>A</b>,<b>D</b>) 2 M, (<b>B</b>,<b>E</b>) 10 M, and (<b>C</b>,<b>F</b>) 20 M group. ALP expression was observed in the osteogenic cells in the new bone matrix. TRAP expression was not specific in all groups (magnification 100×, scale bar = 100 µm).</p>
Full article ">
15 pages, 2466 KiB  
Article
Antimicrobial Activity of Gentamicin-Loaded Biocomposites Synthesized through Inverse Vulcanization from Soybean and Sunflower Oils
by Ana S. Farioli, María V. Martinez, Cesar A. Barbero, Diego F. Acevedo and Edith I. Yslas
Sustain. Chem. 2024, 5(3), 229-243; https://doi.org/10.3390/suschem5030015 - 1 Aug 2024
Viewed by 383
Abstract
Cross-linked polymers synthesized through inverse vulcanization of unsaturated vegetable oils (biopolymers) were used as matrices for incorporating gentamicin (GEN) to form a biocomposite that can amplify GEN antimicrobial activity against Pseudomonas aeruginosa. Two different biopolymers were synthesized using soybean (PSB) and sunflower [...] Read more.
Cross-linked polymers synthesized through inverse vulcanization of unsaturated vegetable oils (biopolymers) were used as matrices for incorporating gentamicin (GEN) to form a biocomposite that can amplify GEN antimicrobial activity against Pseudomonas aeruginosa. Two different biopolymers were synthesized using soybean (PSB) and sunflower (PSF) oils by inverse vulcanization cross-linked with sulfur in a 1:1 weight ratio. The study involves the synthesis and characterization of these biopolymers using FTIR and SEM as well as measurements of density and hydrophobicity. The results reveal the formation of biopolymers, wherein triglyceride molecules undergo cross-linking with sulfur chains through a reaction with the unsaturated groups present in the oil. Additionally, both polymers exhibit a porous structure and display hydrophobic behavior (contact angle higher than 120°). The biopolymers swell more in GEN solution (PSB 127.7% and PSF 174.4%) than in pure water (PSB 88.7% and PSF 109.1%), likely due to hydrophobic interactions. The kinetics of GEN sorption and release within the biopolymer matrices were investigated. The antibacterial efficacy of the resulting biocomposite was observed through the analysis of inhibition growth halos and the assessment of P. aeruginosa viability. A notable enhancement of the growth inhibition halo of GEN (13.1 ± 1.1 mm) compared to encapsulated GEN (PSF-GEN 21.1 ± 1.3 and PSB-GEN 21.45 ± 1.0 mm) is observed. Also, significant bactericidal activity is observed in PSF-GEN and PSB-GEN as a reduction in the number of colonies (CFU/mL), more than 2 log10 compared to control, PSF, and PSB, highlighting the potential of these biopolymers as effective carriers for gentamicin in combating bacterial infections. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of biopolymers: (<b>A</b>) FTIR-ATR spectra of vegetable oils (top) SB (red) and SF (black) and biopolymers (down) PSB (red) and PSF (black). (<b>B1</b>,<b>B2</b>) SEM images of PSB. (<b>C1</b>,<b>C2</b>) SEM images of PSF. (<b>D</b>,<b>E</b>) Contact angle vs. time of PSB (red) and PSF (black).</p>
Full article ">Figure 2
<p>Swelling kinetics for PSF (black) and PSB (red), dots represent the experimental data and lines represent the first-order kinetic fitting for (<b>A</b>) water and (<b>B</b>) 0.5 mg/mL PBS/GEN solution.</p>
Full article ">Figure 3
<p>Kinetics for PSF (black) and PSB (red), dots represent the experimental data: (<b>A</b>) sorption of 0.5 mg/mL PBS/GEN solution, (<b>B</b>) Liberation from biocomposites loaded with 5 mg<sub>GEN</sub>/mL in PBS solution. Line represents the exponential fitting of the experimental data.</p>
Full article ">Figure 4
<p>Inhibition halo of GEN, PSF, PSB, PSF-GEN, and PSB-GEN against <span class="html-italic">P. aeruginosa</span>.</p>
Full article ">Figure 5
<p>Cell Viability of <span class="html-italic">P. aeruginosa</span> of different treatments. Values represent mean ± SE. Different letters indicate significant statistical differences between treatments <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
13 pages, 4609 KiB  
Article
Crosslinking and Swelling Properties of pH-Responsive Poly(Ethylene Glycol)/Poly(Acrylic Acid) Interpenetrating Polymer Network Hydrogels
by Uijung Hwang, HoYeon Moon, Junyoung Park and Hyun Wook Jung
Polymers 2024, 16(15), 2149; https://doi.org/10.3390/polym16152149 - 29 Jul 2024
Viewed by 335
Abstract
This study investigates the crosslinking dynamics and swelling properties of pH-responsive poly(ethylene glycol) (PEG)/poly(acrylic acid) (PAA) interpenetrating polymer network (IPN) hydrogels. These hydrogels feature denser crosslinked networks compared to PEG single network (SN) hydrogels. Fabrication involved a two-step UV curing process: First, forming [...] Read more.
This study investigates the crosslinking dynamics and swelling properties of pH-responsive poly(ethylene glycol) (PEG)/poly(acrylic acid) (PAA) interpenetrating polymer network (IPN) hydrogels. These hydrogels feature denser crosslinked networks compared to PEG single network (SN) hydrogels. Fabrication involved a two-step UV curing process: First, forming PEG-SN hydrogels using poly(ethylene glycol) diacrylate (PEGDA) through UV-induced free radical polymerization and crosslinking reactions, then immersing them in PAA solutions with two different molar ratios of acrylic acid (AA) monomer and poly(ethylene glycol) dimethacrylate (PEGDMA) crosslinker. A subsequent UV curing step created PAA networks within the pre-fabricated PEG hydrogels. The incorporation of AA with ionizable functional groups imparted pH sensitivity to the hydrogels, allowing the swelling ratio to respond to environmental pH changes. Rheological analysis showed that PEG/PAA IPN hydrogels had a higher storage modulus (G′) than PEG-SN hydrogels, with PEG/PAA-IPN5 exhibiting the highest modulus. Thermal analysis via thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) indicated increased thermal stability for PEG/PAA-IPN5 compared to PEG/PAA-IPN1, due to higher crosslinking density from increased PEGDMA content. Consistent with the storage modulus trend, PEG/PAA-IPN hydrogels demonstrated superior mechanical properties compared to PEG-SN hydrogels. The tighter network structure led to reduced water uptake and a higher gel modulus in swollen IPN hydrogels, attributed to the increased density of active network strands. Below the pKa (4.3) of acrylic acid, hydrogen bonds between PEG and PAA chains caused the IPN hydrogels to contract. Above the pKa, ionization of PAA chains induced electrostatic repulsion and osmotic forces, increasing water absorption. Adjusting the crosslinking density of the PAA network enabled fine-tuning of the IPN hydrogels’ properties, allowing comprehensive comparison of single network and IPN characteristics. Full article
(This article belongs to the Special Issue Hydrogels for Biomedical and Structural Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of PEGDA, AA, and PEGDMA.</p>
Full article ">Figure 2
<p>Fabrication process steps for PEG/PAA IPN hydrogels.</p>
Full article ">Figure 3
<p>Raman spectra of hydrogels before and after UV irradiation during the fabrication of PEG/PAA IPN hydrogels.</p>
Full article ">Figure 4
<p>Real-time curing behaviors of PEG-SN, PEG/PAA-IPN1, and PEG/PAA-IPN5.</p>
Full article ">Figure 5
<p>TGA thermograms of PEG-SN, PEG/PAA-IPN1, and PEG/PAA-IPN5.</p>
Full article ">Figure 6
<p>DSC thermograms of PEG-SN, PEG/PAA-IPN1, and PEG/PAA-IPN5.</p>
Full article ">Figure 7
<p>(<b>a</b>) Penetration depth and (<b>b</b>) residual depth profiles of PEG-SN, PEG/PAA-IPN1, and PEG/PAA-IPN5 obtained by NST.</p>
Full article ">Figure 8
<p>AFM images of residual depth profiles of (<b>a</b>) PEG-SN and (<b>b</b>) PEG/PAA-IPN1 at a 0.5 mm scanned position.</p>
Full article ">Figure 9
<p>Swelling ratios as a function of pH for PEG-SN, PEG/PAA-IPN1, and PEG/PAA-IPN5.</p>
Full article ">Figure 10
<p>Moduli obtained from (<b>a</b>) shear tests and (<b>b</b>) compressive tests of swollen hydrogels at different pH levels.</p>
Full article ">
11 pages, 3882 KiB  
Article
Mullite-Fibers-Reinforced Bagasse Cellulose Aerogels with Excellent Mechanical, Flame Retardant, and Thermal Insulation Properties
by Shuang Wang, Miao Sun, Junyi Lv, Jianming Gu, Qing Xu, Yage Li, Xin Zhang, Hongjuan Duan and Shaoping Li
Materials 2024, 17(15), 3737; https://doi.org/10.3390/ma17153737 - 28 Jul 2024
Viewed by 384
Abstract
Cellulose aerogels are considered as ideal thermal insulation materials owing to their excellent properties such as a low density, high porosity, and low thermal conductivity. However, they still suffer from poor mechanical properties and low flame retardancy. In this study, mullite-fibers-reinforced bagasse cellulose [...] Read more.
Cellulose aerogels are considered as ideal thermal insulation materials owing to their excellent properties such as a low density, high porosity, and low thermal conductivity. However, they still suffer from poor mechanical properties and low flame retardancy. In this study, mullite-fibers-reinforced bagasse cellulose (Mubce) aerogels are designed using bagasse cellulose as the raw material, mullite fibers as the reinforcing agent, glutaraldehyde as the cross-linking agent, and chitosan as the additive. The resulted Mubce aerogels exhibit a low density of 0.085 g/cm3, a high porosity of 93.2%, a low thermal conductivity of 0.0276 W/(m∙K), superior mechanical performances, and an enhanced flame retardancy. The present work offers a novel and straightforward strategy for creating high-performance aerogels, aiming to broaden the application of cellulose aerogels in thermal insulation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) SEM image of raw MF. (<b>b</b>) Schematic illustration of the fabrication process of Mubce aerogels. (<b>c</b>) SEM image of as-prepared Mubce-150 aerogels (MF are encapsulated within the BCE, as marked by the yellow arrow, while others are nestled between layers of BCE, as pointed out by the blue arrow). The optical image of (<b>d</b>) Mubce-150 aerogels with different shapes and (<b>e</b>) the photo of the aerogel standing on a petal. And (<b>f</b>) pore size distribution curve of Mubce aerogels obtained by mercury intrusion porosimetry.</p>
Full article ">Figure 2
<p>(<b>a</b>) FTIR spectra of MF, BCE aerogels, and Mubce-150 aerogels; C 1s survey spectra of (<b>b</b>) BCE and (<b>c</b>) Mubce-150 aerogels. N 1s survey spectra of (<b>d</b>) BCE and (<b>e</b>) Mubce-150 aerogels; (<b>f</b>) XRD pattern of BCE and Mubce-150 aerogels; SEM images of (<b>g</b>) BCE aerogels, (<b>h</b>) Mubce-50 aerogels, (<b>i</b>) Mubce-100 aerogels, (<b>j</b>) Mubce-150 aerogels, and (<b>k</b>) Mubce-200 aerogels. And (<b>l</b>) schematic diagram of enhanced mechanical properties of BCE aerogels.</p>
Full article ">Figure 3
<p>(<b>a</b>) Stress–strain curves of BCE and Mubce aerogels. (<b>b</b>) The corresponding compressive strength and density. (<b>c</b>) Photographs of Mubce-150 aerogel withstanding 1561 times its own weight. (<b>d</b>) Stress–strain curves of Mubce-150 aerogels at different compressive strains. Fatigue test of (<b>e</b>) Mubce-150 aerogels and (<b>f</b>) BCE aerogels under 40% compressive strain.</p>
Full article ">Figure 4
<p>(<b>a</b>) Thermal conductivity of Mubce-150 aerogels at different temperatures. (<b>b</b>) Comparison of the volume density and thermal conductivity of Mubce aerogels with other reported aerogels [<a href="#B24-materials-17-03737" class="html-bibr">24</a>,<a href="#B31-materials-17-03737" class="html-bibr">31</a>,<a href="#B32-materials-17-03737" class="html-bibr">32</a>,<a href="#B33-materials-17-03737" class="html-bibr">33</a>,<a href="#B34-materials-17-03737" class="html-bibr">34</a>,<a href="#B35-materials-17-03737" class="html-bibr">35</a>]. (<b>c</b>) Temperature variation curve of the upper surface of the samples heated at 160 °C for 60 min and infrared images of the samples at different heating times on a heating stage. Optical photographs of (<b>d</b>) BCE and (<b>e</b>) Mubce-150 aerogels heated by butane torch flame (aerogel specimen sizes are 90 × 90 × 15 mm<sup>3</sup>). (<b>f</b>) Optical image of Mubce-150 aerogels under butane flame for 45s. SEM iages of Mubce-150 aerogels: (<b>g</b>) area far away from the butane torch flame, (<b>h</b>) carbonized region not directly in contact with the butane torch flame, and (<b>i</b>) ablation region in direct contact with the butane torch flame. And (<b>j</b>) Heat Release Rate (HRR), (<b>k</b>) Total Heat Release (THR), and (<b>l</b>) Mass Loss Rate (MLR) curves of BCE and Mubce-150 aerogels.</p>
Full article ">
17 pages, 4913 KiB  
Article
Comparative Study of the Foaming Behavior of Ethylene–Vinyl Acetate Copolymer Foams Fabricated Using Chemical and Physical Foaming Processes
by Yaozong Li, Junjie Jiang, Hanyi Huang, Zelin Wang, Liang Wang, Bichi Chen and Wentao Zhai
Materials 2024, 17(15), 3719; https://doi.org/10.3390/ma17153719 - 27 Jul 2024
Viewed by 514
Abstract
Ethylene–vinyl acetate copolymer (EVA), a crucial elastomeric resin, finds extensive application in the footwear industry. Conventional chemical foaming agents, including azodicarbonamide and 4,4′-oxybis(benzenesulfonyl hydrazide), have been identified as environmentally problematic. Hence, this study explores the potential of physical foaming of EVA using supercritical [...] Read more.
Ethylene–vinyl acetate copolymer (EVA), a crucial elastomeric resin, finds extensive application in the footwear industry. Conventional chemical foaming agents, including azodicarbonamide and 4,4′-oxybis(benzenesulfonyl hydrazide), have been identified as environmentally problematic. Hence, this study explores the potential of physical foaming of EVA using supercritical nitrogen as a sustainable alternative, garnering considerable interest in both academia and industry. The EVA formulations and processing parameters were optimized and EVA foams with densities between 0.15 and 0.25 g/cm3 were produced. Key findings demonstrate that physical foaming not only reduces environmental impact but also enhances product quality by a uniform cell structure with small cell size (50–100 μm), a wide foaming temperature window (120–180 °C), and lower energy consumption. The research further elucidates the mechanisms of cell nucleation and growth within the crosslinked EVA network, highlighting the critical role of blowing agent dispersion and localized crosslinking around nucleated cells in defining the foam’s cellular morphology. These findings offer valuable insights for producing EVA foams with a more controllable cellular structure, utilizing physical foaming techniques. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the EVA foaming process. (<b>a</b>) Chemical foaming; (<b>b</b>) physical foaming.</p>
Full article ">Figure 2
<p>The vulcanization characteristics and the kinetics analysis of EVA at different temperatures. (<b>a</b>) Dynamic vulcanization curve of odorless BIPB at different dosage, (<b>b</b>,<b>c</b>) were the vulcanization rate of EVA-BIPB where ln (<span class="html-italic">M</span><sub>H</sub> − <span class="html-italic">M</span><sub>t</sub>) varies linearly with time (<b>d</b>) analysis at specified crosslinking degrees by plotting ln K − 1/T and performing fitting.</p>
Full article ">Figure 3
<p>The gel content of EVA formulations. (<span class="html-fig-inline" id="materials-17-03719-i001"><img alt="Materials 17 03719 i001" src="/materials/materials-17-03719/article_deploy/html/images/materials-17-03719-i001.png"/></span>) solid crosslinked EVA sheet used for physical foaming; (<span class="html-fig-inline" id="materials-17-03719-i002"><img alt="Materials 17 03719 i002" src="/materials/materials-17-03719/article_deploy/html/images/materials-17-03719-i002.png"/></span>) EVA foams prepared using chemical foaming.</p>
Full article ">Figure 4
<p>Thermal behavior of raw EVA resin and the crosslinked EVA samples. (<b>a</b>) second heating curve and (<b>b</b>) cooling curve.</p>
Full article ">Figure 5
<p>Desorption of N<sub>2</sub> in E-P-0.5B after saturation for 30 min at different N<sub>2</sub> pressures. The saturation temperature was 120 °C (<b>a</b>). Plots of gas concentration in E-P-0.5B as a function of saturation time at 120 °C (<b>b</b>).</p>
Full article ">Figure 6
<p>Density of EVA foams prepared using chemical and physical foaming at various foaming temperatures, where the saturation pressure was 15 MPa for the physical foaming.</p>
Full article ">Figure 7
<p>SEM images of EVA foams prepared at different foaming temperatures and BIPB loading contents where the EVA formulations were foamed using the chemical method.</p>
Full article ">Figure 8
<p>SEM images of EVA foams prepared at different foaming temperatures and BIPB loading contents, where the EVA formulations were foamed using the physical method.</p>
Full article ">Figure 9
<p>SEM images of EVA foams prepared using a physical process with the saturation temperature of 120 °C and pressures of 10–20 MPa.</p>
Full article ">Figure 10
<p>Mechanisms involved in cell formation and growth within the cross-linked networks during the chemical and physical foaming methods. The pre-foam stage in chemical (<b>a<sub>1</sub></b>), the simultaneous occurrence of cross-linking reactions and foaming agent decomposition during chemical foaming (<b>a<sub>2</sub></b>), and a schematic representation of the structure obtained (<b>a<sub>3</sub></b>); the pre-foam stage in physical foaming (<b>b<sub>1</sub></b>), the nucleation process of bubble formation during saturation (<b>b<sub>2</sub></b>), a schematic diagram depicting the foam structure obtained (<b>b<sub>3</sub></b>).</p>
Full article ">
Back to TopTop