Nothing Special   »   [go: up one dir, main page]

Previous Issue
Volume 10, November
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 

Gels, Volume 10, Issue 12 (December 2024) – 25 articles

  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
17 pages, 12382 KiB  
Article
Microwave Irradiation-Assisted Synthesis of Anisotropic Crown Ether-Grafted Bamboo Pulp Aerogel as a Chelating Agent for Selective Adsorption of Heavy Metals (Mn+)
by Wenxiang Jing, Min Tang, Xiaoyan Lin, Chai Yang, Dongming Lian, Ying Yu and Dongyang Liu
Gels 2024, 10(12), 778; https://doi.org/10.3390/gels10120778 (registering DOI) - 28 Nov 2024
Viewed by 6
Abstract
Crown ether is widely used in water purification because of its ring structure and good selective adsorption of specific heavy metals. However, its high cost and difficulty in recycling limit the purification of heavy metals in water. The anisotropic [2,4]-dibenzo-18-crown-6-modified bamboo pulp aerogel [...] Read more.
Crown ether is widely used in water purification because of its ring structure and good selective adsorption of specific heavy metals. However, its high cost and difficulty in recycling limit the purification of heavy metals in water. The anisotropic [2,4]-dibenzo-18-crown-6-modified bamboo pulp aerogel (DB18C6/PA) is successfully synthesized by microwave irradiation and directional freezing technology. The physical and chemical properties of DB18C6/PA are analyzed by FTIR, XPS, SEM, TEM, TGA, surface area and porosity analyzers. Single or multivariate systems containing Pb2+, Cu2+ and Cd2+ are used as adsorbents. The effects of the DB18C6 addition amount, pH, initial concentration and adsorption temperature on the adsorption of DB18C6/PA are systematically explored. Pseudo-first-order kinetic models, pseudo-second-order kinetic models and the isothermal adsorption models of Langmuir and Freundlich are used to fit the experimental data. The adsorption selectivity is analyzed from the distribution coefficient and the separation factor, and the adsorption mechanism is discussed. The results show that anisotropic DB18C6/PA has the characteristics of 3D directional channels, high porosity (97.67%), large specific surface area (103.7 m2/g), good thermal stability and regeneration (the number of cycles is greater than 5). The surface has a variety of functional groups, including a hydroxyl group, aldehyde group, ether bond, etc. In the single and multivariate systems of Pb2+, Cu2+ and Cd2+, the adsorption process of DB18C6/PA conforms to the pseudo-second-order kinetic model, and the results conform to the Freundlich adsorption isothermal model (a few of them conformed to the Langmuir adsorption isothermal model), indicating that chemical adsorption and physical adsorption are involved in the adsorption process, and the adsorption process is a spontaneous endothermic process. In the single solution system, the maximum adsorption capacities of Pb2+, Cu2+ and Cd2+ by DB18C6/PA are 129.15, 29.85 and 27.89 mg/g, respectively. The adsorption selectivity of DB18C6/PA on Pb2+, Cu2+ and Cd2+ is in the order of Pb2+ >> Cu2+ > Cd2+. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SEM images of DB18C6; (<b>b</b>,<b>c</b>) SEM spectra of DB18C6/PA1-2 obtained by non-directional freezing with different magnification; (<b>d</b>,<b>e</b>) SEM spectra of DB18C6/PA1-2 obtained by directional freezing with different magnifications; (<b>f</b>) TEM of DB18C6/PA1-2 obtained by directional freezing; (<b>g</b>) Digital photos of bamboo pulp aerogel obtained under different preparation conditions; (<b>h</b>) Schematic diagram of DB18C6/PA structure; (<b>i</b>) N<sub>2</sub> adsorption–desorption isotherm curve of DB18C6/PA; (<b>j</b>) Pore size distribution of DB18C6/PA.</p>
Full article ">Figure 2
<p>XPS spectra of DB18C6/PA: (<b>a</b>) full spectrum, (<b>b</b>) C1s spectrum of DB18C6/PA1-1, (<b>c</b>) C1s spectrum of DB18C6/PA1-2, (<b>d</b>) C1s spectrum of DB18C6/PA1-3; (<b>e</b>) TGA and (<b>f</b>) DTG of DB18C6/PA.</p>
Full article ">Figure 3
<p>Effect of DB18C6/PA obtained by non-directional freezing and directional freezing on adsorption of (<b>a</b>) Pb<sup>2+</sup>, (<b>b</b>) Cu<sup>2+</sup>, and (<b>c</b>) Cd<sup>2+</sup>. Effect of DB18C6/PA obtained with different (<b>d</b>) microwave power and (<b>e</b>) reaction time on the adsorption of Pb<sup>2+</sup>, Cu<sup>2+</sup> and Cd<sup>2+</sup>. The effect of (<b>f</b>) pH and (<b>g</b>) initial concentration on the adsorption of Pb<sup>2+</sup>, Cu<sup>2+</sup> and Cd<sup>2+</sup> by DB18C6/PA.</p>
Full article ">Figure 4
<p>The fitted curves of adsorption kinetics models for Pb<sup>2+</sup> in the single, binary and ternary systems of (<b>a</b>) Pb<sup>2+</sup>, (<b>b</b>) Pb<sup>2+</sup>/Cu<sup>2+</sup>, (<b>c</b>) Pb<sup>2+</sup>/Cd<sup>2+</sup>, (<b>d</b>) Pb<sup>2+</sup>/Cu<sup>2+</sup>/Cd<sup>2+</sup>; Cu<sup>2+</sup> in the single, binary and ternary systems of (<b>e</b>) Cu<sup>2+</sup>, (<b>f</b>) Cu<sup>2+</sup>/Pb<sup>2+</sup>, (<b>g</b>) Cu<sup>2+</sup>/Cd<sup>2+</sup>, (<b>h</b>) Cu<sup>2+</sup>/Pb<sup>2+</sup>/Cd<sup>2+</sup>; Cd<sup>2+</sup> in the single, binary and ternary system of (<b>i</b>) Cd<sup>2+</sup>, (<b>j</b>) Cd<sup>2+</sup>/Pb<sup>2+</sup>, (<b>k</b>) Cd<sup>2+</sup>/Cu<sup>2+</sup>, (<b>l</b>) Cd<sup>2+</sup>/Pb<sup>2+</sup>/Cu<sup>2+</sup>, on the DB18C6/PA, at pH = 5 and temperature = 25 °C.</p>
Full article ">Figure 5
<p>The fitted curves of adsorption isothermal models for Pb<sup>2+</sup> in the single, binary and ternary systems of (<b>a</b>) Pb<sup>2+</sup>, (<b>b</b>) Pb<sup>2+</sup>/Cu<sup>2+</sup>, (<b>c</b>) Pb<sup>2+</sup>/Cd<sup>2+</sup>, (<b>d</b>) Pb<sup>2+</sup>/Cu<sup>2+</sup>/Cd<sup>2+</sup>; Cu<sup>2+</sup> in the single, binary and ternary systems of (<b>e</b>) Cu<sup>2+</sup>, (<b>f</b>) Cu<sup>2+</sup>/Pb<sup>2+</sup>, (<b>g</b>) Cu<sup>2+</sup>/Cd<sup>2+</sup>, (<b>h</b>) Cu<sup>2+</sup>/Pb<sup>2+</sup>/Cd<sup>2+</sup>; Cd<sup>2+</sup> in the single, binary and ternary system of (<b>i</b>) Cd<sup>2+</sup>, (<b>j</b>) Cd<sup>2+</sup>/Pb<sup>2+</sup>, (<b>k</b>) Cd<sup>2+</sup>/Cu<sup>2+</sup>, (<b>l</b>) Cd<sup>2+</sup>/Pb<sup>2+</sup>/Cu<sup>2+</sup> on the DB18C6/PA.</p>
Full article ">Figure 6
<p>Adsorption mechanism of DB18C6/PA on Pb<sup>2+</sup>, Cu<sup>2+</sup> and Cd<sup>2+</sup>.</p>
Full article ">Figure 7
<p>Removal efficiency of Pb<sup>2+</sup>, Cu<sup>2+</sup> and Cd<sup>2+</sup> by DB18C6/PA for 5 adsorption–desorption cycles.</p>
Full article ">
15 pages, 4134 KiB  
Article
Nanostructured Hydrogels of Carboxylated Cellulose Nanocrystals Crosslinked by Calcium Ions
by Alexander S. Ospennikov, Yuri M. Chesnokov, Andrey V. Shibaev, Boris V. Lokshin and Olga E. Philippova
Gels 2024, 10(12), 777; https://doi.org/10.3390/gels10120777 (registering DOI) - 28 Nov 2024
Viewed by 64
Abstract
Bio-based eco-friendly cellulose nanocrystals (CNCs) gain an increasing interest for diverse applications. We report the results of an investigation of hydrogels spontaneously formed by the self-assembly of carboxylated CNCs in the presence of CaCl2 using several complementary techniques: rheometry, isothermal titration calorimetry, [...] Read more.
Bio-based eco-friendly cellulose nanocrystals (CNCs) gain an increasing interest for diverse applications. We report the results of an investigation of hydrogels spontaneously formed by the self-assembly of carboxylated CNCs in the presence of CaCl2 using several complementary techniques: rheometry, isothermal titration calorimetry, FTIR-spectroscopy, cryo-electron microscopy, cryo-electron tomography, and polarized optical microscopy. Increasing CaCl2 concentration was shown to induce a strong increase in the storage modulus of CNC hydrogels accompanied by the growth of CNC aggregates included in the network. Comparison of the rheological data at the same ionic strength provided by NaCl and CaCl2 shows much higher dynamic moduli in the presence of CaCl2, which implies that calcium cations not only screen the repulsion between similarly charged nanocrystals favoring their self-assembly, but also crosslink the polyanionic nanocrystals. Crosslinking is endothermic and driven by increasing entropy, which is most likely due to the release of water molecules surrounding the interacting COO and Ca2+ ions. The hydrogels can be easily destroyed by increasing the shear rate because of the alignment of rodlike nanocrystals along the direction of flow and then quickly recover up to 90% of their viscosity in 15 s, when the shear rate is decreased. Full article
(This article belongs to the Special Issue Advances in Cellulose-Based Hydrogels (3rd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Frequency dependencies of storage G′ (filled symbols) and loss G″ (open symbols) moduli for suspensions containing 3 wt% CNCs and different concentrations of CaCl<sub>2</sub>; (<b>b</b>) frequency dependencies of storage G′ (filled) and loss G″ (open) moduli for 3 wt% suspensions of CNCs with 50 mM CaCl<sub>2</sub> (circles) and with 150 mM NaCl (triangles), providing the same ionic strength.</p>
Full article ">Figure 2
<p>Storage modulus G′ (circles) and loss modulus G″ (squares) at the oscillatory frequency of 1 rad/s as a function of CaCl<sub>2</sub> concentration.</p>
Full article ">Figure 3
<p>(<b>a</b>) Viscosity recovery after periodic variation of shear rate (50 s<sup>−1</sup> for 60 s, 0.1 s<sup>−1</sup> for 60 s, etc.) for suspensions containing 3 wt% CNCs and 36 mM (green) and 72 mM (red) of CaCl<sub>2</sub>; (<b>b</b>) fitting of the viscosity recovery with the exponential function for the second cycle of periodic variation of shear rate for 3 wt% CNC suspensions with 36 mM (green) and 72 mM (red) of CaCl<sub>2</sub>. In the formula, η is the apparent viscosity, t is time, τ is the recovery time, a,b are coefficients.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) Cryo-EM images of 3 wt% CNC suspensions before (<b>a</b>,<b>c</b>) and after addition of 50 mM CaCl<sub>2</sub> (<b>b</b>,<b>d</b>) for thinner (<b>a</b>,<b>b</b>) and thicker (<b>c</b>,<b>d</b>) samples. Some domains containing CNCs oriented parallel to each other are marked by ovals. Bundles are marked by yellow arrows, fibrillar-like aggregates are marked by red arrows.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Three-dimensional surface-rendered views of the arrangement of CNCs inside the network before (<b>a</b>) and after addition of 50 mM CaCl<sub>2</sub> (<b>b</b>) obtained from cryo-ET. The bundle is indicated by a yellow arrow, and a fragment of the fibrillar-like aggregate is marked by the red oval.</p>
Full article ">Figure 6
<p>Histograms of distribution of the length (<b>a</b>) and thickness (<b>b</b>) of individual CNCs and their aggregates in 3 wt% CNC suspensions before (green) and after (red) addition of 50 mM CaCl<sub>2</sub>.</p>
Full article ">Figure 7
<p>ITC titration curve of 3 wt% CNC dispersion with CaCl<sub>2</sub> at pH 6.5.</p>
Full article ">Figure 8
<p>(<b>a</b>) ATR-FTIR spectra of CNCs without salt (black) and with 72 mM CaCl<sub>2</sub> (red) in the dried state. The spectra are offset in the <span class="html-italic">y</span>-axis for viewing clarity. (<b>b</b>) ATR-FTIR spectra of 6 wt% suspensions of CNCs without salt (black) and with 72 mM CaCl<sub>2</sub> (red) in water. The spectra are offset in the <span class="html-italic">y</span>-axis for viewing clarity.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>c</b>) Polarized optical microscopy images of CNC suspensions with different concentrations of nanocrystals: 2.5 wt% (<b>a</b>), 3 wt% (<b>b</b>) and 4 wt% (<b>c</b>); (<b>d</b>–<b>f</b>) polarized optical microscopy images of 3 wt% aqueous suspensions of CNCs with different concentrations of added CaCl<sub>2</sub>: 9 mM (<b>d</b>), 18 mM (<b>e</b>) and 36 mM (<b>f</b>).</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>–<b>c</b>) Polarized optical microscopy images of CNC suspensions with different concentrations of nanocrystals: 2.5 wt% (<b>a</b>), 3 wt% (<b>b</b>) and 4 wt% (<b>c</b>); (<b>d</b>–<b>f</b>) polarized optical microscopy images of 3 wt% aqueous suspensions of CNCs with different concentrations of added CaCl<sub>2</sub>: 9 mM (<b>d</b>), 18 mM (<b>e</b>) and 36 mM (<b>f</b>).</p>
Full article ">
17 pages, 3866 KiB  
Article
Preparation and Rheological Evaluation of Thiol–Maleimide/Thiol–Thiol Double Self-Crosslinking Hyaluronic Acid-Based Hydrogels as Dermal Fillers for Aesthetic Medicine
by Chia-Wei Chu, Wei-Jie Cheng, Bang-Yu Wen, Yu-Kai Liang, Ming-Thau Sheu, Ling-Chun Chen and Hong-Liang Lin
Gels 2024, 10(12), 776; https://doi.org/10.3390/gels10120776 (registering DOI) - 28 Nov 2024
Viewed by 124
Abstract
This study presents the development of thiol–maleimide/thiol–thiol double self-crosslinking hyaluronic acid-based (dscHA) hydrogels for use as dermal fillers. Hyaluronic acid with varying degrees of maleimide substitution (10%, 20%, and 30%) was synthesized and characterized, and dscHA hydrogels were fabricated using [...] Read more.
This study presents the development of thiol–maleimide/thiol–thiol double self-crosslinking hyaluronic acid-based (dscHA) hydrogels for use as dermal fillers. Hyaluronic acid with varying degrees of maleimide substitution (10%, 20%, and 30%) was synthesized and characterized, and dscHA hydrogels were fabricated using two molecular weights of four-arm polyethylene glycol (PEG10K/20K)–thiol as crosslinkers. The six resulting dscHA hydrogels demonstrated solid-like behavior with distinct physical and rheological properties. SEM analysis revealed a decrease in porosity with higher crosslinker MW and maleimide substitution. The swelling ratios of the six hydrogels reached equilibrium at approximately 1 h and ranged from 20% to 35%, indicating relatively low swelling. Degradation rates decreased with increasing maleimide substitution, while crosslinker MW had little effect. Higher maleimide substitution also required greater injection force. Elastic modulus (G′) in the linear viscoelastic region increased with maleimide substitution and crosslinker MW, indicating enhanced firmness. All hydrogels displayed similar creep-recovery behavior, showing instantaneous deformation under constant stress. Alternate-step strain tests indicated that all six dscHA hydrogels could maintain elasticity, allowing them to integrate with the surrounding tissue via viscous deformation caused by the stress exerted by changes in facial expression. Ultimately, the connection between the clinical performance of the obtained dscHA hydrogels used as dermal filler and their physicochemical and rheological properties was discussed to aid clinicians in the selection of the most appropriate hydrogel for facial rejuvenation. While these findings are promising, further studies are required to assess irritation, toxicity, and in vivo degradation before clinical use. Overall, it was concluded that all six dscHA hydrogels show promise as dermal fillers for various facial regions. Full article
(This article belongs to the Special Issue Recent Research on Medical Hydrogels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>1H NMR (<b>A</b>) and FTIR spectra (<b>B</b>) of HA and HA-Mal with three different degrees of substitution of maleimide on HA (HM10, HM20, and HM30).</p>
Full article ">Figure 2
<p>Reaction scheme illustrating the formation of <span class="html-italic">dsc</span>HA hydrogels.</p>
Full article ">Figure 3
<p>SEM images of six <span class="html-italic">dsc</span>HA hydrogels (HM10-4SH10K (<b>A</b>), HM20-4SH10K (<b>B</b>), HM30-4SH10K (<b>C</b>), HM10-4SH20K (<b>D</b>), HM20-4SH20K (<b>E</b>), and HM30-4SH20K (<b>F</b>). (Scale bar: 500 µm).</p>
Full article ">Figure 4
<p>The swelling ratio profiles (<b>A</b>) and degradation profiles (<b>B</b>) for HAs with various levels of maleimide substitution and thiol-containing crosslinkers with two different MWs (designated as HM10-4SH10K, HM10-4SH20K, HM20-4SH10K, HM20-4SH20K, HM30-4SH10K, and HM30-4SH20K, respectively).</p>
Full article ">Figure 5
<p>Injection force through a 26 G needle measured for six <span class="html-italic">dsc</span>HA hydrogels (HM10-4SH10K, HM10-4SH20K, HM20-4SH10K, HM20-4SH20K, HM30-4SH10K, and HM30-4SH20K).</p>
Full article ">Figure 6
<p>Rheological evaluation of <span class="html-italic">dsc</span>HA hydrogels. Amplitude sweep (<b>A</b>) and frequency sweep (<b>B</b>) of the six <span class="html-italic">dsc</span>HA hydrogels, showing the linear viscoelastic (LVE) region and gel behavior. Tan δ values (<b>C</b>) of the six <span class="html-italic">dsc</span>HA hydrogels, indicating whether the behavior is elastic-dominant or viscous-dominant.</p>
Full article ">Figure 7
<p>Creep-recovery experiments (constant stress) were performed with an applied shear stress of 5 Pa for 10 min followed by 20 min of recovery (<b>A</b>), and alternate-step strain tests with five repetitions of shear-stress application and relaxation (<b>B</b>) were performed to study the deformation and recovery of the hydrogel network.</p>
Full article ">
18 pages, 9457 KiB  
Article
Novel Injectable Collagen/Glycerol/Pullulan Gel Promotes Osteogenic Differentiation of Mesenchymal Stem Cells and the Repair of Rat Cranial Defects
by Xin Wang, Satoshi Komasa, Yoshiro Tahara, Shihoko Inui, Michiaki Matsumoto and Kenji Maekawa
Gels 2024, 10(12), 775; https://doi.org/10.3390/gels10120775 (registering DOI) - 28 Nov 2024
Viewed by 122
Abstract
Bone tissue engineering is a technique that simulates the bone tissue microenvironment by utilizing cells, tissue scaffolds, and growth factors. The collagen hydrogel is a three-dimensional network bionic material that has properties and structures comparable to those of the extracellular matrix (ECM), making [...] Read more.
Bone tissue engineering is a technique that simulates the bone tissue microenvironment by utilizing cells, tissue scaffolds, and growth factors. The collagen hydrogel is a three-dimensional network bionic material that has properties and structures comparable to those of the extracellular matrix (ECM), making it an ideal scaffold and drug delivery system for tissue engineering. The clinical applications of this material are restricted due to its low mechanical strength. In this investigation, a collagen-based gel (atelocollagen/glycerol/pullulan [Col/Gly/Pul] gel) that is moldable and injectable with high adhesive qualities was created by employing a straightforward technique that involved the introduction of Gly and Pul. This study aimed to characterize the internal morphology and chemical composition of the Col/Gly/Pul gel, as well as to verify its osteogenic properties through in vivo and in vitro experiments. When compared to a standard pure Col hydrogel, this material is more adaptable to the complexity of the local environment of bone defects and the apposition of irregularly shaped flaws due to its greater mechanical strength, injectability, and moldability. Overall, the Col/Gly/Pul gel is an implant that shows great potential for the treatment of complex bone defects and the enhancement of bone regeneration. Full article
(This article belongs to the Special Issue Development of Nanogels/Microgels for Regenerative Medicine)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Fourier-transform infrared spectra: (a) pullulan (Pul), (b) atelocollagen (Col), (c) Col/Gly/Pul gel, and (d) glycerol (Gly). (<b>B</b>,<b>C</b>) Thixotropic studies of Col/Gly/Pul gel and standard liquid. (<b>D</b>) Swelling behavior of the Col/Gly/Pul gel. (<b>E</b>) Amplitude-strain, (<b>F</b>) frequency, and (<b>G</b>) tan δ (<span class="html-italic">G</span>″/<span class="html-italic">G</span>′) of frequency sweeps. LVR, linear viscoelastic region. (<b>H</b>) Images showing the moldability and injectability of the Col/Gly/Pul gel.</p>
Full article ">Figure 2
<p>(<b>A</b>) Scanning electron micrographs, and (<b>B</b>) pore size distribution. (<b>C</b>) The cell viability of rBMSCs initial adherent to the control, Col gel, and Col/Gly/Pul gel. Not significant (ns): <span class="html-italic">p</span> ≥ 0.05; ***: <span class="html-italic">p</span> &lt; 0.001; ****: <span class="html-italic">p</span> &lt; 0.0001. (<b>D</b>) In vitro degradation of the Col/Gly/Pul gel and Col gel.</p>
Full article ">Figure 3
<p>Morphology of rBMSCs attached to the control, Col gel, and Col/Gly/Pul gel after 24 h of incubation. Nuclei were labeled with DAPI (blue), and actin filaments stained with Phalloidin (green). Images obtained at a lower magnification (<b>A</b>) show the number and distribution of cells (scale bars = 100 µm), whereas those obtained at a higher magnification (<b>B</b>) show the connections among cells attached to the different samples (white arrows indicate interactions among cells; scale bars = 50 µm).</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>C</b>) The real-time PCR analysis of specific osteogenesis-related gene expression on the control, Col gel, and Col/Gly/Pul gel. (<b>D</b>) Alkaline phosphatase (ALP) activity and (<b>E</b>) calcium deposition on the control, Col gel, and Col/Gly/Pul gel. Not significant (ns): <span class="html-italic">p</span> ≥ 0.05; *: <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. (<b>F</b>) Alizarin Red staining of the control, Col gel, and Col/Gly/Pul gel (white arrows indicate bone nodules).</p>
Full article ">Figure 5
<p>Effect of the Col/Gly/Pul gel on bone formation in a rat cranial defect model. (<b>A</b>) Treatment protocol for cranial bone defects in rats. Briefly, after separating the periosteum, a 5 mm cylinder-shaped bone defect was prepared on both sides of the sagittal midline of the rat parietal bone and then treated with or without gel. (<b>B</b>) Reconstructed three-dimensional micro-CT images of bone tissues after eight weeks. (<b>C</b>) Bone mineral density (BMD), (<b>D</b>) bone volume-to-total volume ratio (BV/TV), (<b>E</b>) trabecular number (Tb. N), (<b>F</b>) bone surface-to-bone volume ratio (BS/BV), and (<b>G</b>) trabecular separation (Tb. Sp) around the control, Col gel, and Col/Gly/Pul gel after eight weeks. Not significant (ns): <span class="html-italic">p</span> ≥ 0.05; *: <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>Sirius red stain sections of bone tissues after eight weeks around the (<b>A,a,b</b>) Control, (<b>B,c,d</b>) Col gel, and (<b>C,e,f</b>) Col/Gly/Pul gel. Mature, regular fibrous groups were marked by white arrows; immature, disorganized fibrous tissues marked by blue arrows; and the boundary between the old and new bone marked by black arrows.</p>
Full article ">Figure 7
<p>Schematic illustration of the principle of the Col/Gly/Pul gel to promote bone regeneration and repair. (<b>A</b>) Col/Gly/Pul gel provides a scaffold for cell growth. (<b>B</b>) Col/Gly/Pul gel provides mechanical stimuli to the cells and facilitates the diffusion of nutrients. (<b>C</b>) Col/Gly/Pul gel promotes osteogenic differentiation through multiple signaling pathways. (<b>D</b>) Regeneration of new bone tissue with degradation of the Col/Gly/Pul gel. This figure was drawn using Figdraw.</p>
Full article ">Scheme 1
<p>Schematic diagram of the design and preparation of an injectable gel composed of atelocollagen, glycerol, and pullulan (Col/Gly/Pul). (<b>A</b>) Col, Gly, and Pul were stirred at room temperature (23–25 °C) for 5 h. (<b>B</b>) The mixture was frozen at −80 °C for 30 min. (<b>C</b>) The material was dried with a freeze dryer for 12 h. (<b>D</b>) Injectable state of the Col/Gly/Pul gel. (<b>E</b>) Col/Gly/Pul gel before swelling. (<b>F</b>) Network of the swollen Col/Gly/Pul gel.</p>
Full article ">
16 pages, 7199 KiB  
Article
CO2 Foamed Viscoelastic Gel-Based Seawater Fracturing Fluid for High-Temperature Wells
by Jawad Al-Darweesh, Murtada Saleh Aljawad, Muhammad Shahzad Kamal, Mohamed Mahmoud, Shabeeb Alajmei, Prasad B. Karadkar and Bader G. Harbi
Gels 2024, 10(12), 774; https://doi.org/10.3390/gels10120774 (registering DOI) - 27 Nov 2024
Viewed by 258
Abstract
This study investigates the development of a novel CO2-foamed viscoelastic gel-based fracturing fluid to address the challenges of high-temperature formations. The influence of various parameters, including surfactant type and concentration, gas fraction, shear rate, water salinity, temperature, and pressure, on foam [...] Read more.
This study investigates the development of a novel CO2-foamed viscoelastic gel-based fracturing fluid to address the challenges of high-temperature formations. The influence of various parameters, including surfactant type and concentration, gas fraction, shear rate, water salinity, temperature, and pressure, on foam viscosity was systematically explored. Rheological experiments were conducted using a high-pressure/high-temperature (HPHT) rheometer at 150 °C and pressures ranging from 6.89 to 20.68 MPa. To simulate field conditions, synthetic high-salinity water was employed. The thermal stability of the CO2 foam was evaluated at a constant shear rate of 100 1/s for 180 min. Additionally, foamability and foam stability were assessed using an HPHT foam analyzer at 100 °C. The results demonstrate that liquid phase chemistry, experimental conditions, and gas fraction significantly impact foam viscosity. Viscoelastic surfactants achieved a peak foam viscosity of 0.183 Pa·s at a shear rate of 100 1/s and a 70% foam quality, surpassing previous records. At lower foam qualities (≤50%), pressure had a negligible effect on foam viscosity, whereas at higher qualities, it increased viscosity by over 30%. While a slight increase in viscosity was observed with foam qualities between 40% and 60%, a significant enhancement was noted at 65% foam quality. The addition of polymers did not improve foam viscosity. The generation of viscous and stable foams is crucial for effective proppant transport and fracture induction. However, maintaining the thermal stability of CO2 foams with minimal additives remains a significant challenge in the industry. This laboratory study provides valuable insights into the development of stable CO2 foams for stimulating high-temperature wells. Full article
(This article belongs to the Special Issue Gels for Oil and Gas Industry Applications (3rd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Comparative performance of Viscoelastic surfactants for CO<sub>2</sub> foam viscosity at 150 °C, 6.89 MPa, and 70%, (<b>b</b>) thermal foam viscosity at 100 1/s shear rate.</p>
Full article ">Figure 2
<p>The arrangement and configuration of gas bubbles within the CO<sub>2</sub> foam at 100 °C and 6.89 MPa using two viscoelastic systems, surfactant 3 and surfactant 2.</p>
Full article ">Figure 3
<p>(<b>a</b>) Effect of surfactant concentration on CO<sub>2</sub> viscosity at 150 °C, 6.89 MPa, and 70%, (<b>b</b>) thermal foam viscosity at 100 1/s shear rate using three concentrations of surfactant 3 system.</p>
Full article ">Figure 4
<p>(<b>a</b>) Water chemistry affects CO<sub>2</sub> foam viscosity at 150 °C, 6.89 MPa, and 70%, and (<b>b</b>) thermal foam viscosity at 100 1/s shear rate using different water salinities.</p>
Full article ">Figure 5
<p>The arrangement and configuration of gas bubbles within the CO<sub>2</sub> foam at 100 °C and 6.89 MPa for two systems: synthetically prepared seawater (SW) and formation water (FW).</p>
Full article ">Figure 5 Cont.
<p>The arrangement and configuration of gas bubbles within the CO<sub>2</sub> foam at 100 °C and 6.89 MPa for two systems: synthetically prepared seawater (SW) and formation water (FW).</p>
Full article ">Figure 6
<p>(<b>a</b>) Effect of polymers and (<b>b</b>) polymer–mixture on VES CO<sub>2</sub> foam viscosity at 150 °C, 6.89 MPa, 100 1/s, and 70%.</p>
Full article ">Figure 7
<p>The role of elevated pressure and foam quality on CO<sub>2</sub> foam viscosity at 150 °C and a shear rate of 100 1/s.</p>
Full article ">Figure 8
<p>The arrangement and configuration of gas bubbles within the CO<sub>2</sub> at elevated pressure.</p>
Full article ">Figure 8 Cont.
<p>The arrangement and configuration of gas bubbles within the CO<sub>2</sub> at elevated pressure.</p>
Full article ">Figure 9
<p>Water chemistry impact on the (<b>a</b>) average bubble count (mm<sup>−2</sup>) and (<b>b</b>) average bubble radius (μm) at 100 °C and 6.89 MPa.</p>
Full article ">Figure 10
<p>Water chemistry impacts the CO<sub>2</sub> foam volume decay process at 100 °C and 6.89 MPa.</p>
Full article ">Figure 11
<p>Simplified representative of high-pressure, high-temperature foam rheometer.</p>
Full article ">Figure 12
<p>Simplified representative of high-pressure, high-temperature foam analyzer.</p>
Full article ">
20 pages, 3196 KiB  
Article
Reinforcement of Dextran Methacrylate-Based Hydrogel, Semi-IPN, and IPN with Multivalent Crosslinkers
by Luca Paoletti, Gianluca Ferrigno, Nicole Zoratto, Daniela Secci, Chiara Di Meo and Pietro Matricardi
Gels 2024, 10(12), 773; https://doi.org/10.3390/gels10120773 - 27 Nov 2024
Viewed by 209
Abstract
The need for new biomaterials to meet the needs of advanced healthcare therapies is constantly increasing. Polysaccharide-based matrices are considered extremely promising because of their biocompatibility and soft structure; however, their use is limited by their poor mechanical properties. In this light, a [...] Read more.
The need for new biomaterials to meet the needs of advanced healthcare therapies is constantly increasing. Polysaccharide-based matrices are considered extremely promising because of their biocompatibility and soft structure; however, their use is limited by their poor mechanical properties. In this light, a strategy for the reinforcement of dextran-based hydrogels and interpenetrated polymer networks (semi-IPNs and IPNs) is proposed, which will introduce multifunctional crosslinkers that can modify the network crosslinking density. Hydrogels were prepared via dextran methacrylation (DexMa), followed by UV photocrosslinking in the presence of diacrylate (NPGDA), triacrylate (TMPTA), and tetraacrylate (PETA) crosslinkers at different concentrations. The effect of these molecules was also tested on DexMa-gellan semi-IPN (DexMa/Ge) and, later, on IPN (DexMa/CaGe), obtained after solvent exchange with CaCl2 in HEPES and the resulting Ge gelation. Mechanical properties were investigated via rheological and dynamic mechanical analyses to assess the rigidity, resistance, and strength of the systems. Our findings support the use of crosslinkers with different functionality to modulate the properties of polysaccharide-based scaffolds, making them suitable for various biomedical applications. While no significative difference is observed on enriched semi-IPN, a clear improvement is visible on DexMa and DexMa/CaGe systems when TMPTA and NPGDA crosslinker are introduced at higher concentrations, respectively. Full article
(This article belongs to the Special Issue Rheological Properties and Applications of Gel-Based Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of DexMa hydrogels’ (<b>A</b>) and semi-IPNs’ (<b>B</b>) formation, in the presence of diacrylate, triacrylate, and tetraacrylate crosslinkers, after UV photocrosslinking.</p>
Full article ">Figure 2
<p>Dynamic mechanical analysis of DexMa systems. Young’s modulus [E] of DexMa, DexMa-NPGDA, DexMa-TMPTA, and DexMa-PETA hydrogels in a ratio of 10:1 (<b>A</b>), 4:1 (<b>B</b>), and 2:1 (<b>C</b>). Maximum stress registered at the breaking point of DexMa, DexMa-NPGDA, DexMa-TMPTA, and DexMa-PETA hydrogels in a ratio of 10:1 (<b>D</b>), 4:1 (<b>E</b>), and 2:1 (<b>F</b>). Data are expressed as the mean value ± standard deviation; experiments were performed in triplicate (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values were obtained via two-way ANOVA and Dunnett’s multiple comparisons test (**** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Hydrogel rheological analysis, showing the shear storage modulus (G′) at a frequency of 1 Hz, of DexMa, DexMa-NPGDA, DexMa-TMPTA, and DexMa-PETA prepared with a molar ratio of 10:1 (<b>A</b>), 4:1 (<b>B</b>), and 2:1 (<b>C</b>). Tan(delta) loss factor (G″/G′) at 1 Hz of hydrogels obtained with a ratio of 10:1 (<b>D</b>), 4:1 (<b>E</b>), and 2:1 (<b>F</b>). Data are expressed as the mean value ± standard deviation; experiments were performed in triplicate (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values were obtained via two-way ANOVA and Dunnett’s multiple comparisons test (**** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Dynamic mechanical analysis of DexMa/Ge semi-IPNs. Young’s modulus [E] of DexMa/Ge, DexMa-NPGDA/Ge, DexMa-TMPTA/Ge, and DexMa-PETA/Ge semi-IPNs obtained with a 4:1 (<b>A</b>) and 2:1 (<b>B</b>) ratio. Maximum stress registered at the breaking point of DexMa/Ge, DexMa-NPGDA/Ge, DexMa-TMPTA/Ge, and DexMa-PETA/Ge semi-IPNs in a ratio of 4:1 (<b>C</b>) and 2:1 (<b>D</b>). Data are expressed as the mean value ± standard deviation; experiments were performed in triplicate (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values were obtained via two-way ANOVA and Dunnett’s multiple comparisons test (**** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Semi-IPN rheological analysis, showing the shear storage modulus (G′) at a frequency of 1 Hz of DexMa/Ge, DexMa-NPGDA/Ge, DexMa-TMPTA/Ge, and DexMa-PETA/Ge obtained with a molar ratio of 4:1 (<b>A</b>) and 2:1 (<b>B</b>). Tan(delta) loss factor (G″/G′) at 1 Hz of semi-IPN obtained with a ratio of 4:1 (<b>C</b>) and 2:1 (<b>D</b>). Data are expressed as the mean value ± standard deviation; experiments were performed in triplicate (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values were obtained via two-way ANOVA and Dunnett’s multiple comparisons test (**** <span class="html-italic">p</span> &lt; 0.0001; *** <span class="html-italic">p</span> &lt; 0.001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 6
<p>Rheological analysis of DexMa/CaGe, DexMa-NPGDA/CaGe, and DexMa-TMPTA/CaGe IPNs obtained with a 2:1 ratio (<b>A</b>), showing the shear storage modulus (G′) at a frequency of 1 Hz (full bars) and tan(delta) loss factor (G″/G′) at 1 Hz (striped bars). Dynamic mechanical analysis (<b>B</b>): Young’s modulus [E] of DexMa/CaGe, DexMa-NPGDA/CaGe, and DexMa-TMPTA/CaGe IPNs obtained with a 2:1 ratio (full bars) and maximum stress registered at the breaking point (striped bars). Data are expressed as the mean value ± standard deviation; experiments were performed in triplicate (<span class="html-italic">n</span> = 3). <span class="html-italic">p</span> values were obtained via two-way ANOVA and Dunnett’s multiple comparisons test (**** <span class="html-italic">p</span> &lt; 0.0001; ** <span class="html-italic">p</span> &lt; 0.01; * <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
19 pages, 2547 KiB  
Article
The Effect of a Nature-Based Gel on Gingival Inflammation and the Proteomic Profile of Crevicular Fluid: A Randomized Clinical Trial
by Luciene Cristina Figueiredo, Bruno Bueno-Silva, Giovanna Denúncio, Nathalia Freitas Figueiredo, Daniele Ferreira da Cruz, Jamil A. Shibli, Maria Helena R. Borges, Valentim A. R. Barão, Doron Haim, Thabet Asbi and João Gabriel S. Souza
Gels 2024, 10(12), 772; https://doi.org/10.3390/gels10120772 - 27 Nov 2024
Viewed by 234
Abstract
Evidence has shown the clear positive effects of nature-based products on biofilm control and improved gingival health. However, most studies have used in vitro models, have tested single natural components, or have not evaluated proteomic changes after treatment. This double-blind, parallel, randomized, and [...] Read more.
Evidence has shown the clear positive effects of nature-based products on biofilm control and improved gingival health. However, most studies have used in vitro models, have tested single natural components, or have not evaluated proteomic changes after treatment. This double-blind, parallel, randomized, and controlled clinical trial evaluated the benefits of a nature-based gel in controlling gingival inflammation and its effects on the proteomic gingival crevicular fluid (GCF) profile. Gingivitis patients were distributed into the following groups: (1) nature-based gel containing propolis, aloe vera, green tea, cranberry, and calendula (n = 10); (2) control—conventional toothpaste (n = 10). GCF was collected and evaluated by means of liquid chromatography coupled with tandem mass spectrometry (LC–MS/MS). At 3 months, the groups showed similar clinical benefits (p < 0.05). A total of 480 proteins were identified across all groups. In a pooled comparison of both groups at both time points, exclusive proteins were identified in the nature-based gel (78) and the control (21) groups. The exclusive proteins identified for the toothpaste mainly acted in wound healing, and those for the nature-based gel mainly acted on immune system processes. The nature-based gel achieved similar clinical outcomes to conventional toothpaste. However, the nature-based gel markedly changed the proteomic profile of GCF after treatment, showing a profile associated with a host response. Full article
(This article belongs to the Special Issue Designing Gels for Antibacterial and Antiviral Agents)
Show Figures

Figure 1

Figure 1
<p>Proteomic profile of gingival crevicular fluid of patients treated with nature-based gel (DESPLAC<sup>®</sup>) or conventional toothpaste (Oral-B). (<b>A</b>) Total proteins identified for each group and time point (baseline and after 3 months). Proteomic profile was evaluated using liquid chromatography coupled with tandem mass spectrometry. (<b>B</b>) Average LFQ intensity of proteins identified for each group and time point. (<b>C</b>) Venn diagrams comparing the groups and time points.</p>
Full article ">Figure 2
<p>Proteomic profile of gingival crevicular fluid of patients treated with nature-based gel (DESPLAC<sup>®</sup>) or conventional toothpaste (Oral-B) at different time points (BAS—baseline; or 3M—3 months). Heatmap of LFQ intensity of proteins identified for each group and time point.</p>
Full article ">Figure 3
<p>Molecular functions and biological processes mediated by the proteins identified in each group and at each time point. Top 10 molecular functions (<b>A</b>) and biological processes (<b>B</b>) with the highest number of proteins adsorbed onto each substrate. Nature-based gel (DESPLAC<sup>®</sup>) or conventional toothpaste (Oral-B) at different time points (BA—baseline; or 3M—3 months).</p>
Full article ">
19 pages, 5417 KiB  
Article
Effect of Carrot Callus Cells on the Mechanical, Rheological, and Sensory Properties of Hydrogels Based on Xanthan and Konjac Gums
by Elena Günter, Oxana Popeyko, Fedor Vityazev, Natalia Zueva, Inga Velskaya and Sergey Popov
Gels 2024, 10(12), 771; https://doi.org/10.3390/gels10120771 - 27 Nov 2024
Viewed by 190
Abstract
The study aims to develop a plant-based food gel with a unique texture using callus cells and a mixture of xanthan (X) and konjac (K) gums. The effect of encapsulation of carrot callus cells (0.1 and 0.2 g/mL) on properties of X-K hydrogels [...] Read more.
The study aims to develop a plant-based food gel with a unique texture using callus cells and a mixture of xanthan (X) and konjac (K) gums. The effect of encapsulation of carrot callus cells (0.1 and 0.2 g/mL) on properties of X-K hydrogels was studied using the mechanical and rheological analysis with a one-way ANOVA and Student’s t-test used for statistical analysis. Hedonic evaluation and textural features were obtained from 35 volunteers using a nine-point hedonic scale and a 100 mm visual analog scale with the Friedman’s test and the Durbin post hoc test used for statistical analysis. Mechanical hardness, gumminess, and elasticity increased by 1.1–1.3 and 1.1–1.8 times as a result of encapsulation 0.1 and 0.2 g/mL cells, respectively. The addition of cells to the hydrogels resulted in an increase in the complex viscosity, strength, and number of linkages in the gel. The hydrogel samples received identical ratings for overall and consistency liking, as well as taste, aroma, and texture features. However, the callus cell-containing hydrogel had a graininess score that was 82% higher than the callus cell-free hydrogel. The obtained hydrogels based on gums and immobilized carrot callus cells with unique textures may be useful for the development of diverse food textures and the production of innovative functional foods. Full article
(This article belongs to the Special Issue Recent Developments in Food Gels (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Нardness (<b>a</b>), cohesiveness (<b>b</b>), gumminess (<b>c</b>), elasticity (<b>d</b>), springiness (<b>e</b>), and chewiness (<b>f</b>) of hydrogels based on an aqueous mixture of xanthan (X) and konjac (K) gums. The data are presented as the mean ± S.D., <span class="html-italic">n</span> = 12. Different lowercase letters (a, b, and c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the means for different gum concentrations; # <span class="html-italic">p</span> &lt; 0.05 vs. X:K ratio of 1:1, * <span class="html-italic">p</span> &lt; 0.05 vs. X:K ratio of 2:1.</p>
Full article ">Figure 2
<p>Mechanical properties of gum/callus hydrogels based on a 1.0% aqueous mixture of xanthan and konjac gums in different ratios (1:1, 2:1, and 3:1) and 0.1 g/mL carrot callus cells (<b>a</b>). Cell-free gum hydrogels were used as controls (<b>b</b>). Hardness and elasticity are expressed in units of H and mm, respectively. The data are presented as the mean ± S.D., <span class="html-italic">n</span> = 8. # <span class="html-italic">p</span> &lt; 0.05 vs. X:K ratio of 1:1, * <span class="html-italic">p</span> &lt; 0.05 vs. X:K ratio of 2:1, ** <span class="html-italic">p</span> &lt; 0.05 vs. corresponding experimental characteristics in (<b>a</b>).</p>
Full article ">Figure 3
<p>Digital images of cell-free gum hydrogels (X2K1-J) (<b>a</b>) and gum/callus hydrogels based on carrot juice, a 1.0% mixture of xanthan and konjac gums in a 2:1 ratio, and carrot callus cells (0.2 g/mL) (X2K1-0.2DC-J) (<b>b</b>).</p>
Full article ">Figure 4
<p>The effect of carrot callus cell concentration (0.1 and 0.2 g/mL) on the mechanical properties of hydrogels based on carrot juice and a 1.0% mixture of xanthan and konjac gums in a 2:1 ratio. Hardness and elasticity are expressed in units of N and mm, respectively. The data are presented as the mean ± S.D., <span class="html-italic">n</span> = 40. Different lowercase letters (a, b, and c) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the means for different callus cell concentrations. Cell-free gum hydrogels (X2K1-J) were used as controls.</p>
Full article ">Figure 5
<p>Mechanical properties of hydrogels based on a 1.0% mixture of xanthan and konjac gums in a 1:2 ratio, 0.2 g/mL carrot callus cells, and carrot juice. Hardness and elasticity are expressed in units of H and mm, respectively. The data are presented as the mean ± S.D., <span class="html-italic">n</span> = 12. Different lowercase letters (a and b) indicate significant differences (<span class="html-italic">p</span> &lt; 0.05) between the means. # <span class="html-italic">p</span> &lt; 0.05 vs. X2K1-J (X:K ratio of 2:1) in <a href="#gels-10-00771-f004" class="html-fig">Figure 4</a>, * <span class="html-italic">p</span> &lt; 0.05 vs. X2K1-0.2DC-J (X:K ratio of 2:1) in <a href="#gels-10-00771-f004" class="html-fig">Figure 4</a>. Cell-free gum hydrogels (X1K2-J) were used as controls.</p>
Full article ">Figure 6
<p>Rheological properties of hydrogels based on 0.1 and 0.2 g/mL carrot callus cells, carrot juice, and a 1.0% mixture of xanthan and konjac gums in a 2:1 (<b>a</b>,<b>c</b>) and 1:2 (<b>b</b>,<b>d</b>) ratio. Storage modulus (G′, filled symbols) and loss modulus (G″, empty symbols) are represented as a function of shear strain (<b>a</b>,<b>b</b>) or frequency (<b>c</b>,<b>d</b>). Cell-free gum hydrogels (X2K1-J, X1K2-J) were used as controls.</p>
Full article ">Figure 7
<p>Complex viscosity as a function of frequency of hydrogels based on 0.1 and 0.2 g/mL carrot callus cells, carrot juice, and a 1.0% mixture of xanthan and konjac gums in a 2:1 (<b>a</b>) and 1:2 (<b>b</b>) ratio. Cell-free gum hydrogels (X2K1-J, X1K2-J) were used as controls.</p>
Full article ">
20 pages, 1987 KiB  
Article
3D Printed Bigel: A Novel Delivery System for Cannabidiol-Rich Hemp Extract
by Anna Gościniak, Filip Kocaj, Anna Stasiłowicz-Krzemień, Marcin Szymański, Tomasz M. Karpiński and Judyta Cielecka-Piontek
Gels 2024, 10(12), 770; https://doi.org/10.3390/gels10120770 - 26 Nov 2024
Viewed by 466
Abstract
The therapeutic potential of Cannabis sativa L. extract has gained significant attention due to its diverse medical applications. Sublingual administration remains a common delivery method of cannabinoids; however, challenges often arise due to the inconvenient form of the extract and its taste. To [...] Read more.
The therapeutic potential of Cannabis sativa L. extract has gained significant attention due to its diverse medical applications. Sublingual administration remains a common delivery method of cannabinoids; however, challenges often arise due to the inconvenient form of the extract and its taste. To address these issues, a novel bigel formulation was developed, combining water and oil phases to enhance stability and bioavailability. This formulation incorporates a cannabidiol-rich hemp extract, hyaluronic acid for its moisturizing properties, and a taste-masking agent to improve patient compliance and comfort. Using a standardized hemp extract rich in cannabinoids and a well-characterized terpene profile, the printability of the bigels was evaluated through 3D printing technology. A printout with known cannabidiol (CBD) and cannabidiolic acid (CBDA) content of 11.613 mg ± 0.192 of CBD and 4.732 mg ± 0.280 of CBDA in the printout was obtained. In addition, the release profile of CBD and CBDA was evaluated to determine the delivery efficiency of the active ingredient—dissolved active ingredient levels ranged from 74.84% ± 0.50 to 80.87% ± 3.20 for CBD and from 80.84 ± 1.33 to 98.31 ± 1.70 for CBDA depending on the formulation. Rheological studies were conducted to evaluate the viscosity of the bigels under varying temperature conditions, ensuring their stability and usability. Findings suggest that this 3D-printed bigel formulation could significantly enhance the delivery of cannabis extracts, offering a more convenient and effective therapeutic option for patients. This research underscores the importance of innovation in cannabinoid therapies and paves the way for further advancements in personalized medicine. Full article
(This article belongs to the Special Issue Hydrogel for Sustained Delivery of Therapeutic Agents (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Results of the inverted tube test (numbers 1–4 correspond to bigels 1–4).</p>
Full article ">Figure 2
<p>Viscosity of bigels as a function of temperature and velocity (RPM), temperature-dependent viscosity of bigel 1–4 (<b>a</b>,<b>b</b>), viscosity at different velocities (RPM) and temperatures (27 °C, 30 °C, 35 °C)—bigel 1 (<b>c</b>), bigel 2 (<b>d</b>), bigel 3 (<b>e</b>), bigel 4 (<b>f</b>).</p>
Full article ">Figure 3
<p>CBDA and CBDA release curves from bigel formulations 1–4.</p>
Full article ">Figure 4
<p>Star-shaped bigel after printing.</p>
Full article ">
14 pages, 9261 KiB  
Article
Stretchable and Shape-Transformable Organohydrogel with Gallium Mesh Frame
by Mincheol Lee, Youngjin Choi, Young Min Bae, Seonghyeon Nam and Kiyoung Shin
Gels 2024, 10(12), 769; https://doi.org/10.3390/gels10120769 - 26 Nov 2024
Viewed by 264
Abstract
Shape-memory materials are widely utilized in biomedical devices and tissue engineering, particularly for their ability to undergo predefined shape changes in response to external stimuli. In this study, a shape-transformable organohydrogel was developed by incorporating a gallium mesh into a polyacrylamide/alginate/glycerol matrix. The [...] Read more.
Shape-memory materials are widely utilized in biomedical devices and tissue engineering, particularly for their ability to undergo predefined shape changes in response to external stimuli. In this study, a shape-transformable organohydrogel was developed by incorporating a gallium mesh into a polyacrylamide/alginate/glycerol matrix. The gallium mesh, which transitions between solid and liquid states at moderate temperatures (~29.8 °C), enhanced the hydrogel’s mechanical properties and enabled shape-memory functionality. The composite organohydrogel exhibited a high elastic modulus of ~900 kPa in the solid gallium state and ~30 kPa in the liquid gallium state, enabling reversible deformation and structural stability. Glycerol improved the hydrogel’s moisture retention, maintaining stretchability and repeated heating and cooling cycles. After multiple cycles of the shape-changing process, the organohydrogel retained its mechanical integrity, achieving shape-fixation and recovery ratios of ~96% and 95%, respectively. This combination of shape-memory functionality, stretchability, and mechanical stability makes this organohydrogel highly suitable for applications in flexible electronics, soft robotics, and biomedical devices, where adaptability and shape retention are essential. Full article
(This article belongs to the Section Gel Processing and Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Overall schematic of the shape-transformable organohydrogel. (<b>a</b>) The structure of the shape-transformable organohydrogel. (<b>b</b>) The shape transformation cycle of the organohydrogel illustrating the transition between the original and deformed states.</p>
Full article ">Figure 2
<p>Effect of glycerol concentration on the anti-drying properties of organohydrogels. (<b>a</b>) Visual comparison of organohydrogel samples with varying glycerol concentrations (0–50 wt% of aqueous solution) before and after a drying process. The top row shows the original organohydrogel samples, while the bottom row shows the samples after drying. (<b>b</b>) Normalized weight loss of organohydrogels over time in a 60 °C convection oven.</p>
Full article ">Figure 3
<p>Mechanical properties of organohydrogels with varying glycerol concentrations after a drying process. (<b>a</b>) Stress–strain curves of organohydrogels with varying glycerol concentration before drying. Stress–strain curves of the organohydrogel (<b>b</b>) without glycerol and (<b>c</b>) with glycerol after the drying period. (<b>d</b>) Stress relaxation tests under 100% elongation of the organohydrogels.</p>
Full article ">Figure 4
<p>Fabrication process of the shape-transformable organohydrogel.</p>
Full article ">Figure 5
<p>Mechanical properties of the shape-transformable organohydrogel in different states. (<b>a</b>) Microscopic images showing the gallium-reinforced organohydrogel: top view (left) and cross section (right). Red dotted box indicates the layer of the gallium mesh frame. (<b>b</b>) Stress–strain curves comparing the modulus of the organohydrogel in its initial state (black), heated state (red), and transformed state after stretching (blue) and the organohydrogel without gallium frame (green). (<b>c</b>) Elastic modulus of the shape-transformable organohydrogel in different states.</p>
Full article ">Figure 6
<p>Shape transformation and fixation process of the shape-transformable organohydrogel. (<b>a</b>) Organohydrogel without (top) and with gallium mesh frame (bottom). (<b>b</b>) Top view of the organohydrogel with gallium mesh frame (<b>c</b>) Overlay image of sequential deformation of the organohydrogel with gallium mesh frame under heating. Red arrow indicates the moving direction. (<b>d</b>) Image showing the organohydrogel in its fixed state after cooling, with the shape retained from the previous deformation.</p>
Full article ">Figure 7
<p>Bending angle fixation and recovery of the shape-transformable organohydrogel. (<b>a</b>) Schematic representation of the bending angle measurement for shape fixation and recovery. Maximum bending angle, including the maximum bending angle (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>m</mi> <mi>a</mi> <mi>x</mi> <mtext> </mtext> </mrow> </msub> </mrow> </semantics></math> = 180°) and the recovered angle after reheating (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math>). (<b>b</b>) Organohydrogel bent to a maximum angle and cooled in the mold to retain the fixed shape. (<b>c</b>) Organohydrogel returned to its original shape after reheating (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>θ</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math> = 0°).</p>
Full article ">Figure 8
<p>Fixed shapes of the shape-transformable organohydrogel at various bending angles. (<b>a</b>) Organohydrogel sample positioned in an acrylic mold for shape fixation (<b>b</b>–<b>f</b>) Organohydrogel fixed at bending angles of approximately 30°, 60°, 90°, 120° and 150°, respectively, demonstrating the organohydrogel’s ability to retain a range of deformed angles.</p>
Full article ">Figure 9
<p>Step-by-step process for the stretching fixation test of the shape-transformable organohydrogel. (<b>a</b>) Initial length (<span class="html-italic">L<sub>i</sub></span>) measurement of the organohydrogel sample before stretching. (<b>b</b>) Positioning of the heated sample in the stretching device. (<b>c</b>) Sample stretched to the deformed length (<span class="html-italic">L<sub>d</sub></span>) under 100% stretching. (<b>d</b>) Fixed length (<span class="html-italic">L<sub>f</sub></span>) measurement after cooling to maintain the deformed shape. (<b>e</b>) Recovered length (<span class="html-italic">L<sub>r</sub></span>) measurement after reheating, illustrating the organohydrogel’s shape recovery capability.</p>
Full article ">Figure 10
<p>Multiple-cycle stretching fixation and recovery test results for the shape-transformable organohydrogel. (<b>a</b>) Graph showing the shape fixation and recovery ratios over multiple cycles. (<b>b,c</b>) Microscopic images of the gallium frame within the organohydrogel after multiple cycles.</p>
Full article ">Figure 11
<p>Shape transformation and fixation process of the shape-transformable organohydrogel. (<b>a</b>–<b>d</b>) folding and recovery process of the shape-transformable organohydrogel. (<b>e</b>–<b>g</b>) One-dimensional stretching and shape fixation of the shape-transformable organohydrogel. (<b>h</b>) Thermal imaging of the shape recovery process. (<b>i</b>–<b>l</b>) Two-dimensional stretching and shape recovery of the shape-transformable organohydrogel.</p>
Full article ">
37 pages, 9715 KiB  
Review
Molecular Design and Nanoarchitectonics of Inorganic–Organic Hybrid Sol–Gel Systems for Antifouling Coatings
by Markus Bös, Ludwig Gabler, Willi Max Leopold, Max Steudel, Mareike Weigel and Konstantin Kraushaar
Gels 2024, 10(12), 768; https://doi.org/10.3390/gels10120768 - 25 Nov 2024
Viewed by 364
Abstract
Environmental protection, especially fouling protection, is a very topical and wide-ranging issue. This review explores the development, molecular design, and nanoarchitectonics of sol–gel-based hybrid coatings for antifouling applications. These coatings combine inorganic and organic materials, offering enhanced stability and adaptability, making them ideal [...] Read more.
Environmental protection, especially fouling protection, is a very topical and wide-ranging issue. This review explores the development, molecular design, and nanoarchitectonics of sol–gel-based hybrid coatings for antifouling applications. These coatings combine inorganic and organic materials, offering enhanced stability and adaptability, making them ideal for protecting surfaces from fouling. This review covers key antifouling strategies from the past decade, including biocidal additives, fouling resistance, release mechanisms, and surface topological modifications. The sol–gel hybrid systems prevent biofilm formation and organism attachment by leveraging molecular interactions, making them particularly useful in marine environments. Additionally, the study emphasizes the coatings’ environmental benefits, as they offer a potential alternative to traditional toxic antifouling methods. Overall, this research underscores the importance of sol–gel technologies in advancing eco-friendly antifouling solutions. Full article
(This article belongs to the Special Issue Gel Formation and Processing Technologies for Material Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Graphical representation of possible antifouling methods with sol-gel coated surfaces, (<b>b</b>) Graphical outline of the key components in sol–gel chemistry for antifouling coatings.</p>
Full article ">Figure 2
<p>Polysaccharides used by Wanka and Yu et al. to synthesize a hybrid sol–gel network [<a href="#B9-gels-10-00768" class="html-bibr">9</a>].</p>
Full article ">Figure 3
<p>Reaction scheme for the synthesis of LBLHP proposed by Yu et al. [<a href="#B10-gels-10-00768" class="html-bibr">10</a>].</p>
Full article ">Figure 4
<p>Schematic representation of the Baier curve, in which the relative bacterial adhesion is plotted against the surface energy [<a href="#B22-gels-10-00768" class="html-bibr">22</a>].</p>
Full article ">Figure 5
<p>Schematic representation of wetting models (Wenzel left, Cassie–Baxter right).</p>
Full article ">Figure 6
<p>AFM images of 50:50 C8-TES/TEOS coating show a uniform surface (<b>left</b>). The AFM images of 1:49:50 C18-TMS/C8-TES/TEOS coating (<b>right</b>) show a porous surface with 100–300 nm wide pores that are 3–5 nm deep [<a href="#B33-gels-10-00768" class="html-bibr">33</a>].</p>
Full article ">Figure 7
<p>AFM height image of 0.5 mol% PEG in 50:50 TDF/TEOS xerogel (<b>left</b>) and AFM image of the 0.5 mol% PEG in 50:50 TDF/TEOS xerogel in amplitude mode (<b>right</b>) [<a href="#B22-gels-10-00768" class="html-bibr">22</a>].</p>
Full article ">Figure 8
<p>Maximum adhesion force of the BSA to the xerogel surface as a function of the critical surface tension. The measured data show a comparable course to the Baier curve (each value is the average of five repeated measurements) [<a href="#B20-gels-10-00768" class="html-bibr">20</a>].</p>
Full article ">Figure 9
<p>Photos of the coated samples after immersion showing the adhering algae growth (<span class="html-italic">E. crouaniorum</span>) after exposure to shear stress (8 Pa). For the xerogels with aminoalkylsilanes (APTES/TEOS, MAP/TEOS, and DMAP/TEOS), the proportion of bound biomass is significantly higher [<a href="#B31-gels-10-00768" class="html-bibr">31</a>].</p>
Full article ">Figure 10
<p>Relative bacterial adhesion (RBA) to the different xerogels. The higher the telomere content (HP-ZPx), the lower the adhesion of the tested bacterial cultures [<a href="#B81-gels-10-00768" class="html-bibr">81</a>].</p>
Full article ">Figure 11
<p>Plot of the static water contact angles (<b>a</b>) and the surface energy (<b>b</b>) of the natural varnish (RL), the natural varnish with HPSi (RLH10), and the layers to which the telomer (RLH10Sx) was added before and after immersion in water (24 h) [<a href="#B80-gels-10-00768" class="html-bibr">80</a>].</p>
Full article ">Figure 12
<p>Photos of the uncoated (<b>#1</b>) and coated (<b>#2</b>) glass slides after adding chalk powder and applying water to demonstrate the self-cleaning properties of the superhydrophobic sol–gel layer (#2—water rolls off and removes the chalk) [<a href="#B92-gels-10-00768" class="html-bibr">92</a>].</p>
Full article ">Figure 13
<p>Irradiated films with a Bi<sub>2</sub>WO<sub>6</sub> concentration of 50 mM showed no deposition of biomass after 122 days of irradiation [<a href="#B103-gels-10-00768" class="html-bibr">103</a>].</p>
Full article ">Figure 14
<p>The SEM images of the different surfaces show big differences before and after immersion tests for 122 days. In (<b>a</b>), the image of the coated glass before the experiment can be seen. The additional irradiated sample, which can be seen in (<b>b</b>), just shows little differences from (<b>a</b>). The sample with no additional irradiation (<b>c</b>) shows a lot more growth on the surface, while the uncoated glass (<b>d</b>) shows the formation of large biological objects such as diatoms [<a href="#B103-gels-10-00768" class="html-bibr">103</a>].</p>
Full article ">Figure 15
<p>The addition of ZnO nanoparticles improved the resistance of the used substrates against marine fouling according to the visible state of the layers [<a href="#B105-gels-10-00768" class="html-bibr">105</a>].</p>
Full article ">Figure 16
<p>The antifouling properties of the ZnO nanoparticles can be seen from the decrease in the deposited biomass on the coatings [<a href="#B105-gels-10-00768" class="html-bibr">105</a>].</p>
Full article ">Figure 17
<p>Incorporation of only 0.01% of the TiO<sub>2</sub>–nanoparticles synthesized by sol–gel synthesis led to complete killing of the tested bacteria [<a href="#B106-gels-10-00768" class="html-bibr">106</a>].</p>
Full article ">Figure 18
<p>Example of a selenium-containing dendrimer used for antifouling coatings by Detty et al. [<a href="#B104-gels-10-00768" class="html-bibr">104</a>].</p>
Full article ">
16 pages, 3002 KiB  
Article
In Vitro Characterization of Human Cell Sources in Collagen Type I Gel Scaffold for Meniscus Tissue Engineering
by Barbara Canciani, Nicolò Rossi, Elena Arrigoni, Riccardo Giorgino, Mirko Sergio, Lucia Aidos, Mauro Di Giancamillo, Valentina Rafaela Herrera Millar, Giuseppe M. Peretti, Alessia Di Giancamillo and Laura Mangiavini
Gels 2024, 10(12), 767; https://doi.org/10.3390/gels10120767 - 25 Nov 2024
Viewed by 322
Abstract
Strategies to repair the meniscus have achieved limited success; thus, a cell-based therapy combined with an appropriate biocompatible scaffold could be an interesting alternative to overcome this issue. The aim of this project is to analyze different cell populations and a collagen gel [...] Read more.
Strategies to repair the meniscus have achieved limited success; thus, a cell-based therapy combined with an appropriate biocompatible scaffold could be an interesting alternative to overcome this issue. The aim of this project is to analyze different cell populations and a collagen gel scaffold as a potential source for meniscus tissue engineering applications. Dermal fibroblasts (DFs) and mesenchymal stem cells (MSCs) isolated from adipose tissue (ASCs) or bone marrow (BMSCs) were analyzed. Two different fibro-chondrogenic media, M1 and M2, were tested, and qualitative and quantitative analyses were performed. Significant increases in glycosaminoglycans (GAGs) production and in fibro-cartilaginous marker expression were observed in MSCs in the presence of M1 medium. In addition, both ASCs and BMSCs cultured in M1 medium were used in association with the collagen hydrogel (MSCs-SCF) for the development of an in vitro meniscal-like tissue. Significant up-regulation in GAGs production and in the expression of aggrecan, collagen type I, and collagen type II was observed in BMSCs-SCF. This study improves knowledge of the potential of combining undifferentiated MSCs with a collagen gel as a new tissue engineering strategy for meniscus repair. Full article
(This article belongs to the Special Issue Recent Research on Medical Hydrogels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>): Microscopic images showing the fusiform morphology of the three cell populations (MSCs, ASCs, and DFs) cultured under standard conditions. (<b>B</b>): Graph illustrating the doubling time (DT) of the three cell populations in culture, showing that ASCs and DFs have a lower DT compared to BMSCs. (<b>C</b>): Graph representing the total number of cells obtained after one month of culture for each cell population. ASCs and DFs show higher proliferation compared to BMSCs. (<b>D</b>): Graph showing the clonogenic capacity (CFU-F) of the three cell populations. Each cell population maintained good clonogenic capacity. Scalebar: 20 µm.</p>
Full article ">Figure 2
<p>(<b>A</b>): Graph showing the induction of lipid vacuoles in MSCs after 14 days of insulin treatment. ASCs and BMSCs showed a significant increase in lipid production, while DFs showed no induction. (<b>B</b>) (<b>Upper Panel</b>): Graph representing alkaline phosphatase (ALP) activity in MSCs and DFs induced toward the osteogenic lineage. A significant increase in ALP activity was observed in ASCs and BMSCs compared to DFs. (<b>B</b>) (<b>Middle Panel</b>): Graph showing collagen production in osteo-induced cells, with significant increases in all cell populations. (<b>B</b>) (<b>Lower Panel</b>): Graph representing calcium deposition in osteo-induced cells, showing a significant increase for ASCs, BMSCs, and DFs compared to controls. **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>(<b>A</b>): Graph showing GAG production in MSCs and DFs cultured in pellet and treated with M1 and M2 media. A significant increase in GAGs is observed in the presence of M1 medium for ASCs and BMSCs. (<b>B</b>): Histological image showing Safranin O staining in ASCs, BMSCs, and DFs pellets treated with SFM, M1, and M2 medium. Scalebar: 200 µm. **: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>C</b>): Graphs showing gene expression of Aggrecan, Collagen type I, and Collagen type II in cell populations treated with SFM, M1, and M2 medium. *: <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>(<b>A</b>) (<b>Left Panel</b>): Graph illustrating GAG production in differentiated ASCs-SCF, with no significant increases compared to controls (SFM). (<b>A</b>) (<b>Right Panel</b>): Graph showing the increase in GAG production in differentiated BMSCs-SCF in the presence of M1 medium for 7, 14, and 21 days, compared to controls (SFM). (<b>B</b>): Histological images showing Safranin O staining in constructs generated with ASCs and BMSCs. The image highlights extracellular matrix accumulation and a more compact structure in the differentiated constructs, compared to SFM controls. Scalebar: 200 µm. *: <span class="html-italic">p</span> &lt; 0.05; **: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>(<b>A</b>): Graph representing the upregulation of Aggrecan expression in differentiated MSCs (ASCs and BMSCs) in the presence of M1 medium, with an increase over time. (<b>B</b>,<b>C</b>) (<b>Left Panel</b>): Graphs showing Collagen type I and II expression in ASCs, with no induction for Collagen type I and a different trend for Collagen type II. (<b>B</b>,<b>C</b>) (<b>Right Panel</b>): Graphs showing Collagen type I and II expression in BMSCs in the presence of M1 medium, with a significant increase compared to controls. *: <span class="html-italic">p</span> &lt; 0.05; ***: <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>Graphs showing the effect of the collagen sponge on undifferentiated cell populations (SFM ASCs- and BMSCs-SCF). All markers evaluated for fibro-chondrogenic differentiation show a significant increase in SFM bioconstructs maintained for 7, 14, and 21 days compared to day 0, suggesting the inductive potential of the collagen sponge on MSCs. §: <span class="html-italic">p</span> &lt; 0.05; §§: <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
19 pages, 3572 KiB  
Article
The Effect of Cellulose Nanocrystals on the Molecular Organization, Thermomechanical, and Shape Memory Properties of Gelatin-Matrix Composite Films
by Cristina Padilla, Marzena Pępczyńska, Cristian Vizueta, Franck Quero, Paulo Díaz-Calderón, William Macnaughtan, Tim Foster and Javier Enrione
Gels 2024, 10(12), 766; https://doi.org/10.3390/gels10120766 - 25 Nov 2024
Viewed by 371
Abstract
Gelatin is a natural hydrocolloid with excellent film-forming properties, high processability, and tremendous potential in the field of edible coatings and food packaging. However, its reinforcing by materials such as cellulose nanocrystals (CNC) is often necessary to improve its mechanical behavior, including shape [...] Read more.
Gelatin is a natural hydrocolloid with excellent film-forming properties, high processability, and tremendous potential in the field of edible coatings and food packaging. However, its reinforcing by materials such as cellulose nanocrystals (CNC) is often necessary to improve its mechanical behavior, including shape memory properties. Since the interaction between these polymers is complex and its mechanism still remains unclear, this work aimed to study the effect of low concentrations of CNC (2, 6, and 10 weight%) on the molecular organization, thermomechanical, and shape memory properties in mammalian gelatin-based composite films at low moisture content (~10 weight% dry base). The results showed that the presence of CNCs (with type I and type II crystals) interfered with the formation of the gelatin triple helix, with a decrease from 21.7% crystallinity to 12% in samples with 10% CNC but increasing the overall crystallinity (from 21.7% to 22.6% in samples with 10% CNC), which produced a decrease in the water monolayer in the composites. These changes in crystallinity also impacted significantly their mechanical properties, with higher E’ values (from 1 × 104 to 1.3 × 104 Pa at 20 °C) and improved thermal stability at higher CNC content. Additionally, the evaluation of their shape memory properties indicated that while molecular interactions between the two components occur, CNCs negatively impacted the magnitude and kinetics of the shape recovery of the composites (more particularly at 10 weight% CNC, reducing shape recovery from 90% to 70%) by reducing the netting point associated with the lower crystallinity of the gelatin. We believe that our results contribute in elucidating the interactions of gelatin–CNC composites at various structural levels and highlights that even though CNC acts as a reinforcement material on gelatin matrices, their interaction are complex and do not imply synergism in their properties. Further investigation is, however, needed to understand CNC–gelatin interfacial interactions with the aim of modulating their interactions depending on their desired application. Full article
(This article belongs to the Special Issue Design and Development of Gelatin-Based Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The powder X-ray diffraction pattern showing the characteristic diffraction planes of CNCs.</p>
Full article ">Figure 2
<p>FT-IR spectra of gelatin, CNC, and gelatin–CNC composite films equilibrated to 33% RH, ~10 wt.% water content d.b. The highlighted absorption peak located at 1055 cm<sup>−1</sup> is related to the vibrational motions of the C-O stretching of primary alcohols that belong to the molecular structure of cellulose.</p>
Full article ">Figure 3
<p>Gelatin’s and CNCs’ surface distribution at the surface of gelatin–CNC composite films equilibrated to 33% RH, ~10wt.% water content d.b. Raman images show gelatin’s (green) and CNCs’ (red) distribution at the surface of composite films. (<b>A</b>) BG 2 wt.% CNC; (<b>B</b>) BG 6 wt.% CNC; (<b>C</b>) BG 10 wt.% CNC.</p>
Full article ">Figure 4
<p>X-ray diffraction patterns of gelatin, CNC, and gelatin–CNC composite films equilibrated to 33% RH, ~10 wt.% water content d.b with various CNC weight percentages.</p>
Full article ">Figure 5
<p>Sorption isotherms at 20 °C of gelatin, CNC, and gelatin–CNC composite films with different CNC weight fractions. Lines in the graphs represent the fitting to the experimental data using the GAB equation. The inset shows isotherms at low moisture content.</p>
Full article ">Figure 6
<p>DSC thermograms (first heating scans) of gelatin and gelatin–CNC composite films with different CNC weight fractions equilibrated under 33% RH, ~10wt.% water content d.b. Black arrows indicate the glass transition temperature (Tg) shift with the presence of CNC.</p>
Full article ">Figure 7
<p>Storage modulus (E’) of gelatin and gelatin–CNC composite films with different CNC weight fractions equilibrated at RH 33% at 20 and 50 °C. Data from 4 independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, *** <span class="html-italic">p</span> &lt; 0.0005.</p>
Full article ">Figure 8
<p>Thermograms of gelatin, CNC (inset), and gelatin–CNC composite films with various CNC weight percentages equilibrated at RH 33%.</p>
Full article ">Figure 9
<p>The percentage of the shape recovery of gelatin and gelatin–CNC composite films with different CNC weight fractions at 30 °C as a function of time.</p>
Full article ">
21 pages, 10455 KiB  
Article
Experimental Evaluation of a Recrosslinkable CO2-Resistant Micro-Sized Preformed Particle Gel for CO2 Sweep Efficiency Improvement in Reservoirs with Super-K Channels
by Adel Alotibi, Tao Song, Ali Al Brahim, Baojun Bai and Thomas Schuman
Gels 2024, 10(12), 765; https://doi.org/10.3390/gels10120765 - 24 Nov 2024
Viewed by 248
Abstract
A recrosslinkable CO2-resistant branched preformed particle gel (CO2-BRPPG) was developed for controlling CO2 injection conformance, particularly in reservoirs with super-permeable channels. Previous work focused on a millimeter-sized CO2-BRPPG in open fractures, but its performance in high-permeability [...] Read more.
A recrosslinkable CO2-resistant branched preformed particle gel (CO2-BRPPG) was developed for controlling CO2 injection conformance, particularly in reservoirs with super-permeable channels. Previous work focused on a millimeter-sized CO2-BRPPG in open fractures, but its performance in high-permeability channels with pore throat networks remained unexplored. This study used a sandpack model to evaluate a micro-sized CO2-BRPPG under varying conditions of salinity, gel concentration, and pH. At ambient conditions, the equilibrium swelling ratio (ESR) of the gel reached 76 times its original size. This ratio decreased with increasing salinity but remained stable at low pH values, demonstrating the gel’s resilience in acidic environments. Rheological tests revealed shear-thinning behavior, with gel strength improving as salinity increased (the storage modulus rose from 113 Pa in 1% NaCl to 145 Pa in 10% NaCl). Injectivity tests showed that lower gel concentrations reduced the injection pressure, offering flexibility in deep injection treatments. Gels with higher swelling ratios had lower injection pressures due to increased strength and reduced deformability. The gel maintained stable plugging performance during two water-alternating-CO2 cycles, but a decline was observed in the third cycle. It also demonstrated a high CO2 breakthrough pressure of 177 psi in high salinity conditions (10% NaCl). The permeability reduction for water and CO2 was influenced by gel concentration and salinity, with higher salinity increasing the permeability reduction and higher gel concentrations decreasing it. These findings underscore the effectiveness of the CO2-BRPPG in improving CO2 sweep efficiency and managing CO2 sequestration in reservoirs with high permeability. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) A micro-size RPPG is injected to fill the pore space within the high-permeability channel. (<b>B</b>) RPPG is associated with forming a bulk, continuous-phase gel.</p>
Full article ">Figure 2
<p>(<b>a</b>) Effect of salinity on particle size. (<b>b</b>) Effect of pH on particle size.</p>
Full article ">Figure 3
<p>Viscosity as a function of the shear rate of the CO<sub>2</sub>-BRPPG suspension at different suspension concentrations.</p>
Full article ">Figure 4
<p>(<b>a</b>) The linear viscoelastic region of the CO<sub>2</sub>-BRPPG demonstrates consistent modulus values within this range. (<b>b</b>) The effect of salinity on gel strength.</p>
Full article ">Figure 5
<p>(<b>a</b>) Swelling ratio of the CO<sub>2</sub>-BRPPG before and after exposure to CO<sub>2</sub> at pressures of 500 psi, 850 psi (dense phase), and 1200 psi (supercritical CO<sub>2</sub>). (<b>b</b>) Effect of salinity on the swelling ratio of the CO<sub>2</sub>-BRPPG under 850 psi CO<sub>2</sub> pressure.</p>
Full article ">Figure 6
<p>Set of pictures showing the CO<sub>2</sub>-BRPPG after exposure to CO<sub>2</sub> at 850 psi for different salinities.</p>
Full article ">Figure 7
<p>SEM picture for the gel before (<b>a</b>) and after (<b>b</b>) exposure to supercritical CO<sub>2</sub> for 3 days.</p>
Full article ">Figure 8
<p>Profile of the injection pressure on the CO<sub>2</sub>-BRPPG microgel at different salinity and concentration values.</p>
Full article ">Figure 9
<p>(<b>a</b>) Relationship of CO<sub>2</sub>-BRPPG concentration and stable injection pressure gradient. (<b>b</b>) The CO<sub>2</sub>-BRPPG suspension salinity and stable injection pressure gradient.</p>
Full article ">Figure 10
<p>Differential pressure profile of water-alternating-CO<sub>2</sub> injection in a sandpack treated with 3000 ppm CO<sub>2</sub>-BRPPG.</p>
Full article ">Figure 11
<p>Differential pressure of the first CO<sub>2</sub> injection cycle.</p>
Full article ">Figure 12
<p>Differential pressure of the first brine injection cycle.</p>
Full article ">Figure 13
<p>(<b>a</b>,<b>b</b>): CO<sub>2</sub> breakthrough pressure for different gel concentrations and salinity conditions.</p>
Full article ">Figure 14
<p>The residual resistance factor for both CO<sub>2</sub> and water at different flow rates (1, 1.25, 1.5, and 1.75 mL/min) and gel concentrations (1500 ppm, 3000 ppm, 5000 ppm, and 7000 ppm).</p>
Full article ">Figure 15
<p>The relationship between gel concentration and RDPR reduction for CO<sub>2</sub> and water.</p>
Full article ">Figure 16
<p>The residual resistance factor for both CO<sub>2</sub> and water at different flow rates (1, 1.25, 1.5, and 1.75 mL/min) under varying salinity conditions with a constant gel concentration of 5000 ppm.</p>
Full article ">Figure 17
<p>The relationship between NaCl concentration and RDPR for CO<sub>2</sub> and water.</p>
Full article ">Figure 18
<p>(<b>a</b>) shows the CO<sub>2</sub>-BRPPG bulk gel that formed after free-radical polymerization, (<b>b</b>) shows the CO<sub>2</sub>-BRPPG ground to a 170/230 mesh size, (<b>c</b>) shows a microscopic picture of dry CO<sub>2</sub>-BRPPG, and (<b>d</b>) shows a swelling particle gel.</p>
Full article ">Figure 19
<p>High-pressure vessel and loaded sample.</p>
Full article ">Figure 20
<p>Schematic of the experimental setup for CO<sub>2</sub>/brine flooding.</p>
Full article ">
15 pages, 4134 KiB  
Article
Enhanced Ammonia Capture for Adsorption Heat Pumps Using a Salt-Embedded COF Aerogel Composite
by Hiluf T. Fissaha and Duckjong Kim
Gels 2024, 10(12), 764; https://doi.org/10.3390/gels10120764 - 24 Nov 2024
Viewed by 313
Abstract
Adsorption heat pumps (AHPs) have garnered significant attention due to their efficient use of low-grade thermal energy, eco-friendly nature, and cost-effectiveness. However, a significant challenge lies in developing adsorbent materials that can achieve a high uptake capacity, rapid adsorption rates, and efficient reversible [...] Read more.
Adsorption heat pumps (AHPs) have garnered significant attention due to their efficient use of low-grade thermal energy, eco-friendly nature, and cost-effectiveness. However, a significant challenge lies in developing adsorbent materials that can achieve a high uptake capacity, rapid adsorption rates, and efficient reversible release of refrigerants, such as ammonia (NH3). Herein, we developed and synthesized a novel salt-embedded covalent organic framework (COF) composite material designed for enhanced NH3 capture. This material was prepared by encapsulating sodium bromide (NaBr) within a porous and densely functionalized sulfonic acid-based COF. The COF was synthesized through a Schiff base (imine) condensation reaction, providing a robust platform for effective NaBr impregnation. The COF-based aerogel composite powder was investigated for its potential in ammonia-based AHPs, benefiting from both the porous, highly functionalized COF structure and the strong NH3 affinity of the impregnated NaBr. The composite adsorbent demonstrates an impressive NH3 adsorption capacity, adsorption rate, and stability. The exceptional NH3 adsorption performance of the COF-based aerogel composite powder is primarily attributed to the uniformly dispersed NaBr within the COF, the coordination of NH3 molecules with Na+ ions, and the hydrogen bonding interaction between NH3 and Br- ions. These findings highlight the potential of the salt-embedded COF composite for use in NH3-based AHPs, gas separation, and other related applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Field emission scanning electron microscopy (FE-SEM) analysis of (<b>a</b>) NaBr@SACOF-80 and (<b>b</b>) SACOF at both low and high magnification, along with corresponding energy dispersive spectroscopy (EDS) analysis of selected elements.</p>
Full article ">Figure 2
<p>BET analysis of NaBr@SACOF-80 and SACOF. (<b>a</b>) N<sub>2</sub> adsorption isotherms, (<b>b</b>) pore size distribution, and (<b>c</b>) surface area, pore size, and pore volume.</p>
Full article ">Figure 3
<p>FTIR analysis of NaBr@SCCOF-80 and its intermediates.</p>
Full article ">Figure 4
<p>Structural and thermal analysis of NaBr@SACOF-80 and its intermediates (<b>a</b>) XRD and (<b>b</b>) TGA analysis.</p>
Full article ">Figure 5
<p>Adsorption characteristics of NaBr@SACOF composites, SACOF, and pure NaBr: (<b>a</b>) amount of adsorbed NH<sub>3</sub> (%) between the temperature range of 80 °C to 20 °C; (<b>b</b>) time profile for NaBr@SACOF-80, SACOF, and pure NaBr; and the fitted non-linear kinetic model (N.B: better fitted model only shown here).</p>
Full article ">Figure 6
<p>Adsorption characteristics: (<b>a</b>) adsorption capacity of SACOF as a function of NaBr weight percentage and adsorption temperature and (<b>b</b>) cyclic NH<sub>3</sub> adsorption capacity of NaBr@SACOF-80.</p>
Full article ">Figure 7
<p>Adsorption properties of the as-prepared NaBr@SACOF compared with the literature [<a href="#B9-gels-10-00764" class="html-bibr">9</a>,<a href="#B10-gels-10-00764" class="html-bibr">10</a>,<a href="#B24-gels-10-00764" class="html-bibr">24</a>,<a href="#B25-gels-10-00764" class="html-bibr">25</a>,<a href="#B28-gels-10-00764" class="html-bibr">28</a>,<a href="#B29-gels-10-00764" class="html-bibr">29</a>].</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic layout of our customized high-pressure ammonia sorption analyzer, (<b>b</b>) actual picture of the experimental setup, and (<b>c</b>) high-pressure ammonia reactor with pressure and temperature sensors [<a href="#B29-gels-10-00764" class="html-bibr">29</a>].</p>
Full article ">Scheme 1
<p>Synthesis scheme of SACOF and NaBr@SACOF.</p>
Full article ">
12 pages, 1837 KiB  
Article
Modifying the Resistant Starch Content and the Retrogradation Characteristics of Potato Starch Through High-Dose Gamma Irradiation
by Zhangchi Peng, Xuwei Wang, Zhijie Liu, Liang Zhang, Linrun Cheng, Jiahao Nia, Youming Zuo, Xiaoli Shu and Dianxing Wu
Gels 2024, 10(12), 763; https://doi.org/10.3390/gels10120763 - 24 Nov 2024
Viewed by 368
Abstract
Potato starch is widely utilized in the food industry. Gamma irradiation is a cost-effective and environmentally friendly method for starch modification. Nevertheless, there is a scarcity of comprehensive and consistent knowledge regarding the physicochemical characteristics of high-dose gamma-irradiated potato starch, retrogradation properties in [...] Read more.
Potato starch is widely utilized in the food industry. Gamma irradiation is a cost-effective and environmentally friendly method for starch modification. Nevertheless, there is a scarcity of comprehensive and consistent knowledge regarding the physicochemical characteristics of high-dose gamma-irradiated potato starch, retrogradation properties in particular. In this study, potato starch was exposed to gamma rays at doses of 0, 30, 60, 90, and 120 kGy. Various physicochemical properties, including retrogradation characteristics, were investigated. Generally, the apparent amylose content (AAC), water absorption, gel viscosity, gel hardness, and gumminess decreased as the doses of gamma irradiation increased. Conversely, the resistant starch (RS), amylose content evaluated by the concanavalin A precipitation method, water solubility, and enthalpy of gelatinization were increased. Additionally, swelling power, crystalline structure, and amylopectin branch chain length distribution either remained stable or exhibited only minor changes. Notably, the degree of retrogradation of potato starches on day 7 was positively correlated with the doses of gamma irradiation. Full article
(This article belongs to the Special Issue Natural Bioactive Compounds and Gels)
Show Figures

Figure 1

Figure 1
<p>Morphology of native and gamma-irradiated potato starches. CK: native starch, 30 kGy, 60 kGy, 90 kGy and 120 kGy indicated the starch irradiated by 30, 60, 90, and 120 kGy gamma irradiation.</p>
Full article ">Figure 2
<p>X-ray diffraction (XRD) patterns (<b>a</b>) and Fourier transformed infrared (FTIR) spectra (<b>b</b>) of native and gamma-irradiated potato starches. The typical hexagonal X-ray diffraction peaks are pointed out by arrows.</p>
Full article ">Figure 3
<p>Amylopectin branch chain length distribution of native and gamma-irradiated potato starches (<b>a</b>) and the differential chain length distribution of amylopectin between samples irradiated with 0 Gy and other dosages (<b>b</b>).</p>
Full article ">Figure 4
<p>X-ray diffraction (XRD) patterns of gelatinized native and gamma-irradiated potato starches after 7 days of retrogradation at 4 °C.</p>
Full article ">
30 pages, 3826 KiB  
Review
Exosome-Integrated Hydrogels for Bone Tissue Engineering
by Hee Sook Hwang and Chung-Sung Lee
Gels 2024, 10(12), 762; https://doi.org/10.3390/gels10120762 - 23 Nov 2024
Viewed by 232
Abstract
Exosome-integrated hydrogels represent a promising frontier in bone tissue engineering, leveraging the unique biological properties of exosomes to enhance the regenerative capabilities of hydrogels. Exosomes, as naturally occurring extracellular vesicles, carry a diverse array of bioactive molecules that play critical roles in intercellular [...] Read more.
Exosome-integrated hydrogels represent a promising frontier in bone tissue engineering, leveraging the unique biological properties of exosomes to enhance the regenerative capabilities of hydrogels. Exosomes, as naturally occurring extracellular vesicles, carry a diverse array of bioactive molecules that play critical roles in intercellular communication and tissue regeneration. When combined with hydrogels, these exosomes can be spatiotemporally delivered to target sites, offering a controlled and sustained release of therapeutic agents. This review aims to provide a comprehensive overview of the recent advancements in the development, engineering, and application of exosome-integrated hydrogels for bone tissue engineering, highlighting their potential to overcome current challenges in tissue regeneration. Furthermore, the review explores the mechanistic pathways by which exosomes embedded within hydrogels facilitate bone repair, encompassing the regulation of inflammatory pathways, enhancement of angiogenic processes, and induction of osteogenic differentiation. Finally, the review addresses the existing challenges, such as scalability, reproducibility, and regulatory considerations, while also suggesting future directions for research in this rapidly evolving field. Thus, we hope this review contributes to advancing the development of next-generation biomaterials that synergistically integrate exosome and hydrogel technologies, thereby enhancing the efficacy of bone tissue regeneration. Full article
(This article belongs to the Special Issue Recent Advances in Hydrogels for Biomedical Application (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Exosome-integrated hydrogels for bone tissue engineering. Exosomes, which develop from the endosomal system within cells, are packed with proteins, lipids, and nucleic acids that facilitate robust intercellular communication. Exosomes, typically around 40–100 nm in size, belong to a diverse group of extracellular vesicles, set apart by their phospholipid membrane and specific formation pathway. They are released by a variety of cell types and circulate in bodily fluids, with their molecular composition reflecting the origin and condition of the parent cell. Exosomes can be isolated and fabricated through differential centrifugation, density gradient centrifugation, ultrafiltration, size exclusion chromatography, affinity nanoparticle-based isolation, polymer precipitation, microfluidic technologies, etc. Exosomes contain diverse bioactive molecules such as proteins, nucleic acids, and lipids, with two main types of proteins: general markers (like CD9, CD63, and CD81) and proteins unique to their parent cell. Exosomes can be modified with drug loading, active targeting, and stimuli-responsiveness. Hydrogels have become valuable tools in biomedical applications for exosome delivery. Hydrogels, as water-retentive polymer networks, encapsulate exosomes, improving their retention at target sites and providing controlled release to enhance localized effects. These exosome–hydrogel composites show promise in areas such as bone tissue engineering and treatment. (Created with <a href="https://biorender.com" target="_blank">https://biorender.com</a>).</p>
Full article ">Figure 2
<p>Human umbilical cord mesenchymal stem cell-derived exosomes with hydroxyapatite-embedded hyaluronic acid-alginate hydrogel for bone regeneration. (<b>A</b>) Schematic illustration of human umbilical cord mesenchymal stem cell-derived exosomes with hydroxyapatite-embedded hyaluronic acid-alginate hydrogel. (<b>B</b>) Transmission electron microscopy image of exosomes. (<b>C</b>) Size distribution of exosomes. (<b>D</b>) Western blot analysis of the exosome surface markers. (<b>E</b>) Release profiles of exosomes from the hydrogels with or without hydroxyapatite. * <span class="html-italic">p</span> &lt; 0.05 compared to the control group; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05 compared to the hydrogel group. (<b>F</b>) Confocal microscopy images of hydrogel with or without red fluorescence DiI-labeled exosomes. (<b>G</b>) Reconstructed 3D micro-CT images of the exosome-integrated hydrogels. Red circles indicate the defect area. (<b>H</b>) Quantitative analysis of bone regeneration using bone volume/tissue volume (BV/TV). Reproduced with permission from Yang et al. [<a href="#B164-gels-10-00762" class="html-bibr">164</a>].</p>
Full article ">Figure 3
<p>Exosome (Exo)-encapsulated stem cell-recruitment hydrogel microcarriers for osteoarthritis (OA) treatment. (<b>A</b>) Schematic illustration of Exo-encapsulated stem cell recruitment particles for OA treatment. (<b>a</b>) Hyaluronic acid (HA) and gelatin-based polymer matrix and bioactive components of hydrogel microcarriers. (<b>b</b>) Fabrication process of hydrogel microcarriers. (<b>c</b>) Application of microcarriers for OA treatment. (<b>B</b>–<b>D</b>) Characterization of particles and Exo. (<b>B</b>) Microscopic image of the particles. (<b>C</b>) The enlarged view of (<b>B</b>). (<b>D</b>) The size distribution of the particles. (<b>E</b>,<b>F</b>) H&amp;E and Safranin O-fast green staining results. (<b>E</b>) H&amp;E results after different particle treatments. (<b>F</b>) Safranin O-fast green staining results after the particle treatment. Par = microfluidic electrospray-generated hydrogel particles; Par@Pep = Par modified by SKPPGTSS peptides; Par@Exo = Par loaded with Exo; Par@Pep&amp;Exo = Par@Pep loaded with Exo. Reproduced with permission from Yang et al. [<a href="#B173-gels-10-00762" class="html-bibr">173</a>].</p>
Full article ">
20 pages, 4539 KiB  
Article
Development of Soft Wrinkled Micropatterns on the Surface of 3D-Printed Hydrogel-Based Scaffolds via High-Resolution Digital Light Processing
by Mauricio A. Sarabia-Vallejos, Scarleth Romero De la Fuente, Nicolás A. Cohn-Inostroza, Claudio A. Terraza, Juan Rodríguez-Hernández and Carmen M. González-Henríquez
Gels 2024, 10(12), 761; https://doi.org/10.3390/gels10120761 - 23 Nov 2024
Viewed by 220
Abstract
The preparation of sophisticated hierarchically structured and cytocompatible hydrogel scaffolds is presented. For this purpose, a photosensitive resin was developed, printability was evaluated, and the optimal conditions for 3D printing were investigated. The design and fabrication by additive manufacturing of tailor-made porous scaffolds [...] Read more.
The preparation of sophisticated hierarchically structured and cytocompatible hydrogel scaffolds is presented. For this purpose, a photosensitive resin was developed, printability was evaluated, and the optimal conditions for 3D printing were investigated. The design and fabrication by additive manufacturing of tailor-made porous scaffolds were combined with the formation of surface wrinkled micropatterns. This enabled the combination of micrometer-sized channels (100–200 microns) with microstructured wrinkled surfaces (1–3 μm wavelength). The internal pore structure was found to play a critical role in the mechanical properties. More precisely, the TPMS structure with a zero local curvature appears to be an excellent candidate for maintaining its mechanical resistance to compression stress, thus retaining its structural integrity upon large uniaxial deformations up to 70%. Finally, the washing conditions selected enabled us to produce noncytotoxic materials, as evidenced by experiments using AlamarBlue to follow the metabolic activity of the cells. Full article
Show Figures

Figure 1

Figure 1
<p>Schematical description of the process followed to obtain the wrinkled scaffolds.</p>
Full article ">Figure 2
<p>(<b>a</b>) FT-IR spectra of liquid and polymerized resin for sample 300:50:150, and (<b>b</b>) magnification of vinyl bands (C=C, 1619–1635 cm<sup>−1</sup>, red dashed box) where Gaussian fitting (green line) of the peaks is depicted.</p>
Full article ">Figure 3
<p>STL representation of the CAD models used for test 3D-printed scaffolds: gyroid-TPMS (<b>left</b>), solid (<b>center</b>), and straight cylindrical channels (<b>right</b>).</p>
Full article ">Figure 4
<p>Stress–strain curves from the compressive mechanical tests of the three different structures studied.</p>
Full article ">Figure 5
<p>OM images of the surface wrinkled micropatterns obtained at different vacuum exposure times from 30 min to 180 min.</p>
Full article ">Figure 6
<p>(<b>a</b>) AFM micrographs and FFT image; (<b>b</b>) wavelength (black line) and amplitude (red line) of the wrinkled patterns and roughness of the samples using different times and speed rotations via spin coating, according to Table 6 for sample 300:50:150.</p>
Full article ">Figure 7
<p>FE-SEM micrographs of wrinkled surfaces with different compositions: (<b>a</b>) 500:0:0, (<b>b</b>) 400:100:0, (<b>c</b>) 400:0:100, (<b>d</b>) 300:100:100, (<b>e</b>) 300:50:150 of PEGDA:DMAEMA:AAm.</p>
Full article ">Figure 8
<p>Cell viability results based on the AlamarBlue assay for various resin compositions at 1, 3, and 7 days of culture. Data are expressed as mean ± standard deviation (<span class="html-italic">n</span> = 3). * <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; n.s., nonsignificant differences.</p>
Full article ">Figure 9
<p>Adhesion of MC3T3-E1 cells on resin surfaces at 24 h of incubation. Bar: 50 µM, magnification: 20×, 40×, and 63×.</p>
Full article ">
15 pages, 4814 KiB  
Article
Development of Variable Charge Cationic Hydrogel Particles with Potential Application in the Removal of Amoxicillin and Sulfamethoxazole from Water
by Francisca L. Aranda, Manuel F. Meléndrez, Mónica A. Pérez, Bernabé L. Rivas, Eduardo D. Pereira and Daniel A. Palacio
Gels 2024, 10(12), 760; https://doi.org/10.3390/gels10120760 - 23 Nov 2024
Viewed by 215
Abstract
Cationic hydrogel particles (CHPs) crosslinked with glutaraldehyde were synthesized and characterized to evaluate their removal capacity for two globally consumed antibiotics: amoxicillin and sulfamethoxazole. The obtained material was characterized by FTIR, SEM, and TGA, confirming effective crosslinking. The optimal working pH was determined [...] Read more.
Cationic hydrogel particles (CHPs) crosslinked with glutaraldehyde were synthesized and characterized to evaluate their removal capacity for two globally consumed antibiotics: amoxicillin and sulfamethoxazole. The obtained material was characterized by FTIR, SEM, and TGA, confirming effective crosslinking. The optimal working pH was determined to be 6.0 for amoxicillin and 4.0 for sulfamethoxazole. Under these conditions, the CHPs achieved over 90.0% removal of amoxicillin after 360 min at room temperature, while sulfamethoxazole removal reached approximately 60.0% after 300 min. Thermodynamic analysis indicated that adsorption occurs through a physisorption process and is endothermic. The ΔH° values of 28.38 kJ mol−1, 12.39 kJ mol−1, and ΔS° 97.19 J mol−1 K−1, and 33.94 J mol−1 K−1 for AMX and SMX, respectively. These results highlight the potential of CHPs as promising materials for the removal of such contaminants from aqueous media. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of the antibiotics sulfamethoxazole (<b>a</b>), amoxicillin (<b>b</b>), and their respective pKa values.</p>
Full article ">Figure 2
<p>FTIR spectra of CHPs.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM-EDS microstructure of CHIPs (Chi-Control and Chi-Glu5%) and size distribution; (<b>b</b>) thermogravimetric analysis of CHIPs (Chi-Control and Chi-Glu5%).</p>
Full article ">Figure 4
<p>(<b>a</b>) CHPs crosslinked with glutaraldehyde at different concentrations subjected to a water absorption process over time and (<b>b</b>) comparison of water absorption at equilibrium time 8 h. * Significant at <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>(<b>a</b>) Identification of the point of zero charge (PZC) of chitosan CHPs; (<b>b</b>) adsorption mediated by electrostatic interactions and hydrogen bonding between the crosslinked chitosan CHPs and the study antibiotics; removal of the antibiotics (<b>c</b>) amoxicillin and (<b>d</b>) sulfamethoxazole at different pH values. ** Significant at <span class="html-italic">p</span> &lt; 0.05, * Significant at <span class="html-italic">p</span> &lt; 0.01, ns = no significant difference.</p>
Full article ">Figure 6
<p>Effect of the ionic strength of monovalent and divalent ions on the removal of the antibiotics (<b>a</b>,<b>c</b>) sulfamethoxazole and (<b>b</b>,<b>d</b>) amoxicillin; effect of the adsorbent dose on the removal of (<b>e</b>) amoxicillin and (<b>f</b>) sulfamethoxazole. Those values that are very significant are not expressed graphically, only those of low significance and those that are not significant are detailed. ** Significant at <span class="html-italic">p</span> &lt; 0.05, * Significant at <span class="html-italic">p</span> &lt; 0.01, ns = no significant difference.</p>
Full article ">Figure 7
<p>Absorption kinetics of the antibiotics amoxicillin (<b>a</b>) and sulfamethoxazole (<b>b</b>) by Chi-Glu CHPs as a function of temperature, (<b>c</b>) template effect.</p>
Full article ">Figure 8
<p>Effect of concentration variation on the adsorption of antibiotics amoxicillin (<b>a</b>) and sulfamethoxazole (<b>b</b>) by Chi-Glu at different temperatures.</p>
Full article ">Figure 9
<p>Process for obtaining cationic hydrogel particles.</p>
Full article ">Figure 10
<p>Calibration curves of the antibiotics (<b>a</b>) sulfamethoxazole and (<b>b</b>) amoxicillin at pH values 3.0–8.0.</p>
Full article ">
17 pages, 11736 KiB  
Article
3D Printing of New Foods Using Cellulose-Based Gels Obtained from Cerotonia siliqua L. Byproducts
by Antoni Capellà, Mónica Umaña, Esperanza Dalmau, Juan A. Cárcel and Antoni Femenia
Gels 2024, 10(12), 759; https://doi.org/10.3390/gels10120759 (registering DOI) - 23 Nov 2024
Viewed by 249
Abstract
Carob pulp is a valuable source of cellulose-rich fraction (CRF) for many food applications. This study aimed to obtain and characterize a CRF derived from carob pulp waste after sugar removal and to evaluate its potential use in the 3D printing of cellulose-rich [...] Read more.
Carob pulp is a valuable source of cellulose-rich fraction (CRF) for many food applications. This study aimed to obtain and characterize a CRF derived from carob pulp waste after sugar removal and to evaluate its potential use in the 3D printing of cellulose-rich foods. Thus, the extraction of the CRF present in carob pulp (by obtaining the alcohol-insoluble residue) was carried out, accounting for nearly 45% dm (dry matter) of this byproduct. The CRF contained about 24% dm of cellulose. The functional properties (swelling capacity, water retention, and fat adsorption) related to this fraction were determined, showing a value of 5.9 mL/g of CRF and 4.0 and 6.5 g/g of CRF, respectively. Different gels were formulated with a total solids content of 15% wm (wet matter), using potato peel flour as a base and partially substituting with CRF (0% to 8% wm). The cellulose-based gels were characterized in terms of viscosity, water distribution (low-field Nuclear Magnetic Resonance), and printability, while the 3D printed samples were assessed for their textural properties. As the percentage of added CRF increased, the viscosity decreased while the water retention increased. Printability improved when small proportions of CRF (2% to 4%) were used, while it deteriorated for higher percentages (6% to 8%). The textural properties (hardness, adhesiveness, cohesiveness, and gumminess) showed significant changes caused by the addition of CRF, with gels containing 3% to 4% CRF exhibiting the most suitable printing values. In summary, this study demonstrates the significant potential of carob cellulose-based gel as an ingredient in the 3D printing of novel fiber-rich foods, contributing to reducing food waste and promoting sustainable practices within the framework of the circular economy. Full article
(This article belongs to the Special Issue Cellulose-Based Gels: Synthesis, Properties, and Applications)
Show Figures

Figure 1

Figure 1
<p>Apparent viscosity of the cellulose-based gels formulated with 0%, 2%, 3%, 4%, 6%, and 8% CRF. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Distributions of the relaxation time after inverse Laplace transform analysis of the LF-NMR spectra of the cellulose-based gels formulated with 0%, 2%, 3%, 4%, 6%, and 8% CRF.</p>
Full article ">Figure 3
<p>Textural properties: hardness (<b>a</b>), adhesiveness (<b>b</b>), cohesiveness (<b>c</b>), and gumminess (<b>d</b>) of the printed gels (shape-stable) formulated with 0%, 2%, 3%, 4%, 6%, and 8% CRF. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Percentages of porosity for the printed samples across the different formulations. Different lowercase letters indicate significant differences (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 5
<p>Photograph of the carob pulp and seeds of the <span class="html-italic">Bugadera</span> variety.</p>
Full article ">
17 pages, 7663 KiB  
Article
Preparation and Characterization of Poly(acrylic acid-co-vinyl imidazole) Hydrogel-Supported Palladium Catalyst for Tsuji–Trost and Suzuki Reactions in Aqueous Media
by Huijun Song, Amatjan Sawut, Rena Simayi and Yuqi Sun
Gels 2024, 10(12), 758; https://doi.org/10.3390/gels10120758 - 23 Nov 2024
Viewed by 292
Abstract
In this study, a novel polyacrylate-co-vinyl imidazole hydrogel-supported palladium (Pd) catalyst (P(AA-co-VI)@Pd) was prepared through heat-initiated polymerization, starting with the formation of a complex between vinyl imidazole and palladium chloride, followed by the addition of 75% neutralized acrylic acid (AA), crosslinking agent, and [...] Read more.
In this study, a novel polyacrylate-co-vinyl imidazole hydrogel-supported palladium (Pd) catalyst (P(AA-co-VI)@Pd) was prepared through heat-initiated polymerization, starting with the formation of a complex between vinyl imidazole and palladium chloride, followed by the addition of 75% neutralized acrylic acid (AA), crosslinking agent, and initiator. The structure and morphology of the catalyst were characterized using ICP-OES, SEM, EDX, Mapping, FT-IR, TGA, XRD, XPS and TEM techniques. It was confirmed that the catalyst exhibited excellent compatibility with water solvent and uniform distribution of Pd. The P(AA-co-VI)@Pd hydrogel catalyst demonstrated remarkable catalytic activity and ease of separation. Notably, in a Tsuji–Trost reaction, employing water as the solvent, it achieved a conversion rate as high as 94% at very low catalyst dosages, indicating its superior catalytic performance. Moreover, after six consecutive cycles, the catalyst maintained good activity and structural stability, highlighting its exceptional reusability and environmental friendliness. Furthermore, the outstanding efficiency of the catalyst was also observed in a Suzuki coupling reaction where both conversion rate and yield reached 100% and 99%, respectively, within just one hour reaction time, thus further validating its universality and efficacy across various chemical reactions. Full article
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>,<b>b</b>) SEM of P(AA-co-VI) hydrogels of 50 µm and 20 µm. (<b>c</b>,<b>d</b>) are SEM of the P(AA-co-VI)@Pd hydrogels.</p>
Full article ">Figure 2
<p>(<b>a</b>) EDS spectrum and (<b>b</b>) Mapping analysis diagram of P(AA-co-VI)@Pd.</p>
Full article ">Figure 3
<p>FT-IR of raw material and the P(AA-co-VI)@Pd.</p>
Full article ">Figure 4
<p>TGA of P(AA-co-VI) and P(AA-co-VI)@Pd.</p>
Full article ">Figure 5
<p>XRD of P(AA-co-VI) and P(AA-co-VI)@Pd.</p>
Full article ">Figure 6
<p>(<b>a</b>) XPS scan image and the high-resolution spectroscopy of C 1s (<b>b</b>), N 1s (<b>c</b>), N 1s (<b>d</b>), C 1s (<b>e</b>), and Pd3d (<b>f</b>) of the catalyst P(AA-co-VI) and P(AA-co-VI)@Pd.</p>
Full article ">Figure 7
<p>The TEM micrographs of P(AA-co-VI)@Pd: (<b>a</b>,<b>b</b>) after 1st time recovered, (<b>c</b>,<b>d</b>) after 6th time recovered.</p>
Full article ">Figure 8
<p>Swelling and mechanical properties of (<b>a</b>,<b>b</b>) P(AA-co-VI) and P(AA-co-VI)@Pd.</p>
Full article ">Figure 9
<p>Recovery of P(AA-co-VI)@Pd in the Tsuji–Trost reaction. Reaction conditions: NaBPh<sub>4</sub> (0.342 g, 1.0 mmol), P(AA-co-VI)@Pd catalyst (10 mg, 0.012 µmol Pd), cinnamyl acetate (0.088 g, 0.5 mmol), and water (5 mL) reflux at 80 °C.</p>
Full article ">Figure 10
<p>Mechanism of the Tsuji–Trost and Suzuki reactions.</p>
Full article ">Figure 11
<p>Chemical structure formula (<b>a</b>) and preparation process (<b>b</b>) of the P(AA-co-VI)@Pd catalyst.</p>
Full article ">
15 pages, 3295 KiB  
Article
Electrostatic Gelatin Nanoparticles for Biotherapeutic Delivery
by Connor Tobo, Avantika Jain, Madhushika Elabada Gamage, Paul Jelliss and Koyal Garg
Gels 2024, 10(12), 757; https://doi.org/10.3390/gels10120757 - 23 Nov 2024
Viewed by 365
Abstract
Biological agents such as extracellular vesicles (EVs) and growth factors, when administered in vivo, often face rapid clearance, limiting their therapeutic potential. To address this challenge and enhance their efficacy, we propose the electrostatic conjugation and sequestration of these agents into gelatin-based biomaterials. [...] Read more.
Biological agents such as extracellular vesicles (EVs) and growth factors, when administered in vivo, often face rapid clearance, limiting their therapeutic potential. To address this challenge and enhance their efficacy, we propose the electrostatic conjugation and sequestration of these agents into gelatin-based biomaterials. In this study, gelatin nanoparticles (GNPs) were synthesized via the nanoprecipitation method, with adjustments to the pH of the gelatin solution (4.0 or 10.0) to introduce either a positive or negative charge to the nanoparticles. The GNPs were characterized using dynamic light scattering (DLS), X-ray diffraction (XRD), Fourier-transform infrared spectroscopy (FTIR), and Transmission electron microscopy (TEM) imaging. Both positively and negatively charged GNPs were confirmed to be endotoxin-free and non-cytotoxic. Mesenchymal stem cell (MSC)-derived EVs exhibited characteristic surface markers and a notable negative charge. Zeta potential measurements validated the electrostatic conjugation of MSC-EVs with positively charged GNPs. Utilizing a transwell culture system, we evaluated the impact of EV-GNP conjugates encapsulated within a gelatin hydrogel on macrophage secretory activity. The results demonstrated the bioactivity of EV-GNP conjugates and their synergistic effect on macrophage secretome over five days of culture. In summary, these findings demonstrate the efficacy of electrostatically coupled biotherapeutics with biomaterials for tissue regeneration applications. Full article
(This article belongs to the Special Issue Hydrogel for Tissue Engineering and Biomedical Therapeutics)
Show Figures

Figure 1

Figure 1
<p>GNP size, morphology, and zeta potential characterization. The size distribution of acidic (<b>A</b>) and alkaline (<b>B</b>) GNPs is shown. TEM images show particle morphology and integrity of acidic (<b>C</b>) and alkaline (<b>D</b>) GNPs (scale bar = 200 nm). (<b>E</b>) Comparison of zeta potential of acidic GNPs vs. alkaline GNPs is presented. Data reported as mean ± standard deviation. A <span class="html-italic">t</span>-test demonstrates statistical significance (*** denotes <span class="html-italic">p</span> &lt; 0.001, n = 3).</p>
Full article ">Figure 2
<p>Powder XRD spectra comparing stock type-A gelatin (<b>A</b>) and acidic GNPs (<b>B</b>). FTIR spectra of both (<b>C</b>,<b>E</b>) type A and B gelatin and (<b>D</b>,<b>F</b>) type A and type B GNPs.</p>
Full article ">Figure 3
<p>LDH Assay assessing cytotoxicity. No statistically significant differences were detected between GNP treatment groups and untreated media control (# denotes significance between all other groups, <span class="html-italic">p</span> &lt; 0.001 (n = 3)).</p>
Full article ">Figure 4
<p>MSC-EV size and surface marker characterization. (<b>A</b>) NTA data shows an average EV particle size range of 50–300 nm (n = 5). NTA for 1X PBS is also shown as negative control (n = 3). (<b>B</b>) An exo-check array reveals the presence of EV-specific markers, confirming the successful isolation of EVs. A marker for cellular contamination (GM130) was negative.</p>
Full article ">Figure 5
<p>(<b>A</b>) Zeta potential measurements comparing acidic GNPs and EVs versus GNP<sup>+</sup>:EV conjugates. From left to right, GNP<sup>+</sup>:EV ratios tested include 400:1, 200:1, and 100:1. Each GNP and conjugate measurement maintained a consistent GNP concentration of 1.0 mg/mL while EV concentration was changed between groups. Different EV concentrations are specified (x-axis). (<b>B</b>) The zeta potential of alkaline GNPs, EVs, and their respectively combined solutions in the same ratios of 400:1, 200:1, and 100:1. Statistical comparisons between measurements of different GNP<sup>+</sup> to EV ratios are not shown for ease of viewing. Data is represented as mean ± standard deviation (n = 3, statistical significance is denoted by * for <span class="html-italic">p</span> &lt; 0.05 between groups for One-way ANOVA).</p>
Full article ">Figure 6
<p>Release of GNP<sup>+</sup>:EV Conjugates and Their Temporal Effects on Bioactivity. (<b>A</b>) IL-6 secretion is significantly reduced in groups treated with GNP<sup>+</sup>:EV conjugates over the course of both three and five days, and it is also reduced on day three for cells cultured with control gels (<b>B</b>) Macrophages treated with EVs and GNP<sup>+</sup>:EV conjugates secreted significantly more VEGF than cells treated with gels containing only GNPs. (<b>C</b>) IGF-1 concentration is not significantly different amongst any groups. (<b>D</b>) No differences in cellular proliferation were noted between groups (n = 4, * denotes <span class="html-italic">p</span> &lt; 0.05, a repeated measures two-way ANOVA was used for analysis).</p>
Full article ">Figure 7
<p>GNP<sup>+</sup>:EV conjugation increases particle size and offers protection for improved retention of biological cargo.</p>
Full article ">Figure 8
<p>GNP synthesis using the nanoprecipitation method and carbodiimide crosslinking prior to resuspension.</p>
Full article ">
15 pages, 6312 KiB  
Article
Environmentally Friendly Nanoporous Polymeric Gels for Sustainable Wastewater Treatment
by Tarek M. Madkour, Rasha E. Elsayed and Rasha A. Azzam
Gels 2024, 10(12), 756; https://doi.org/10.3390/gels10120756 - 22 Nov 2024
Viewed by 305
Abstract
Environmentally friendly nanoporous gels are tailor-designed and employed in the adsorption of toxic organic pollutants in wastewater. To ensure the maximum adsorption of the contaminant molecules by the gels, molecular modeling techniques were used to evaluate the binding affinity between the toxic organic [...] Read more.
Environmentally friendly nanoporous gels are tailor-designed and employed in the adsorption of toxic organic pollutants in wastewater. To ensure the maximum adsorption of the contaminant molecules by the gels, molecular modeling techniques were used to evaluate the binding affinity between the toxic organic contaminants such as methylene blue (MB) and Congo red (CR) and various biopolymers. To generate nanopores in the matrix of the polymeric gels, salt crystals were used as porogen. The pores were then used to accommodate catalytic nickel (Ni0) nanoparticles. Under UV irradiation, the nanoparticles demonstrated the effective adsorption and photocatalytic degradation of both the methylene blue and Congo red dyes, achieving removal efficiencies of up to 90% for MB and 80% for CR. The thermodynamic analysis suggested a spontaneous endothermic dissociative adsorption mechanism, which implies the oxidative catalytic degradation of the dyes. The kinetic modeling suggested a pseudo-second-order model, while the model for intra-particle diffusion revealed that Congo red diffuses faster than methylene blue. MB adsorption followed a Langmuir isotherm, while CR adsorption followed a linear isotherm. The results confirm that dye molecules initially undergo physisorption and subsequent dissociative adsorption. The products of the catalytic degradation of methylene blue continue to be absorbed on the surface of the nanoparticles, while those of Congo red switch to preferential desorption. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Snapshots of the molecular simulations of (<b>a</b>) the polymer blend mixed with MB dye molecules and (<b>b</b>) the polymer blend mixed with CR dye molecules. Each cell has a triclinic lattice with a side length of 19.7064 Å and an angle of 90°.</p>
Full article ">Figure 2
<p>Enthalpy of mixing MB (<b>a</b>) and CR (<b>b</b>) with the different polymeric blends at different dye mole fractions.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM image for the unloaded gel, (<b>b</b>) SEM image for the Ni-loaded gel, (<b>c</b>) TGA for the unloaded gel, and (<b>d</b>) TGA for the Ni-loaded gel.</p>
Full article ">Figure 4
<p>BET isotherms of the unloaded gel (<b>a</b>) and the Ni<sup>0</sup>-loaded gel (<b>b</b>).</p>
Full article ">Figure 5
<p>The uptake capacity and percent removal of methylene blue and Congo red at equilibrium at various initial concentrations of the dye solutions (<b>a</b>), used dose of the adsorbent (<b>b</b>), and operating temperature (<b>c</b>).</p>
Full article ">Figure 6
<p>Plots of van’t Hoff (<b>a</b>) and equilibrium isotherms (<b>b</b>) for MB and CR adsorption.</p>
Full article ">Figure 7
<p>The kinetic plots of the removal of methylene blue and Congo red dyes (<b>a</b>), linear isotherms estimated using the pseudo-first-order model (<b>b</b>), linear isotherms estimated using the pseudo-second-order model (<b>c</b>), and linear isotherms estimated using the intra-particle diffusion model (<b>d</b>).</p>
Full article ">Scheme 1
<p>Schematic diagram for the preparation of in situ impregnated polymeric gels.</p>
Full article ">
20 pages, 1695 KiB  
Review
Recent Advances of Cellulose-Based Hydrogels Combined with Natural Colorants in Smart Food Packaging
by Lan Yang, Qian-Yu Yuan, Ching-Wen Lou, Jia-Horng Lin and Ting-Ting Li
Gels 2024, 10(12), 755; https://doi.org/10.3390/gels10120755 - 21 Nov 2024
Viewed by 332
Abstract
Due to the frequent occurrence of food safety problems in recent years, healthy diets are gradually receiving worldwide attention. Chemical pigments are used in smart food packaging because of their bright colors and high visibility. However, due to shortcomings such as carcinogenicity, people [...] Read more.
Due to the frequent occurrence of food safety problems in recent years, healthy diets are gradually receiving worldwide attention. Chemical pigments are used in smart food packaging because of their bright colors and high visibility. However, due to shortcomings such as carcinogenicity, people are gradually looking for natural pigments to be applied in the field of smart food packaging. In traditional smart food packaging, the indicator and the packaging bag substrate have different degrees of toxicity. Smart food packaging that combines natural colorants and cellulose-based hydrogels is becoming more and more popular with consumers for being natural, non-toxic, environmentally friendly, and renewable. This paper reviews the synthesis methods and characteristics of cellulose-based hydrogels, as well as the common types and characteristics of natural pigments, and discusses the application of natural colorants and cellulose-based hydrogels in food packaging, demonstrating their great potential in smart food packaging. Full article
(This article belongs to the Special Issue Advances in Cellulose-Based Hydrogels (3rd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The repeating unit of cellulose, also known as “cellulobiose”. (<b>b</b>) Repeating units of cellulose derivatives. Substituent “R” for methyl cellulose (MC), hydroxypropyl methyl cellulose (HPMC), ethyl cellulose (EC), hydroxyethyl cellulose (HEC), and sodium carboxymethyl cellulose (NaCMC).</p>
Full article ">Figure 2
<p>Sources of cellulose and applications of CHBs.</p>
Full article ">
25 pages, 8025 KiB  
Review
The Unfulfilled Potential of Synthetic and Biological Hydrogel Membranes in the Treatment of Abdominal Hernias
by Kenigen Manikion, Christodoulos Chrysanthou and Constantinos Voniatis
Gels 2024, 10(12), 754; https://doi.org/10.3390/gels10120754 - 21 Nov 2024
Viewed by 403
Abstract
Hydrogel membranes can offer a cutting-edge solution for abdominal hernia treatment. By combining favorable mechanical parameters, tissue integration, and the potential for targeted drug delivery, hydrogels are a promising alternative therapeutic option. The current review examines the application of hydrogel materials composed of [...] Read more.
Hydrogel membranes can offer a cutting-edge solution for abdominal hernia treatment. By combining favorable mechanical parameters, tissue integration, and the potential for targeted drug delivery, hydrogels are a promising alternative therapeutic option. The current review examines the application of hydrogel materials composed of synthetic and biological polymers, such as polyethylene glycol (PEG), polyvinyl alcohol (PVA), gelatine, and silk fibroin, in the context of hernia repair. Overall, this review highlights the current issues and prospects of hydrogel membranes as viable alternatives to the conventional hernia meshes. The emphasis is placed on the applicability of these hydrogels as components of bilayer systems and standalone materials. According to our research, hydrogel membranes exhibit several advantageous features relevant to hernia repair, such as a controlled inflammatory reaction, tissue integration, anti-adhesive-, and even thermoresponsive properties. Nevertheless, despite significant advancements in material science, the potential of hydrogel membranes seems neglected. Bilayer constructs have not transitioned to clinical trials, whereas standalone membranes seem unreliable due to the lack of comprehensive mechanical characterization and long-term in vivo experiments. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The current literature review PRISMA 2020 flowchart [<a href="#B3-gels-10-00754" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>An umbilical hernia and its treatment options. (Used with permission from the Hernia Center of Southern Carolina, USA).</p>
Full article ">Figure 3
<p>Examples of synthetic meshes. (<b>a</b>) Bard Mesh, (<b>b</b>) Vypro, (<b>c</b>) Prolene V R, (<b>d</b>) Bard V R Soft Mesh, (<b>e</b>) Trelex V, (<b>f</b>) Optilene V R, (<b>g</b>) SurgiPro V R, (<b>h</b>) Parietene V R, (<b>i</b>) Mersilene, and (<b>j</b>) Dynamesh V R IPOM. Used with kind permission from Elsevier [<a href="#B20-gels-10-00754" class="html-bibr">20</a>].</p>
Full article ">Figure 4
<p>An example of a biological mesh: an Acellular Dermal Matrix (Allomend by Allosource, <a href="https://hcp.alloderm.com/" target="_blank">https://hcp.alloderm.com/</a>, accessed on 10 November 2024).</p>
Full article ">Figure 5
<p>Advantages and disadvantages of synthetic and biological meshes.</p>
Full article ">Figure 6
<p>Examples of cross-linking tactics [<a href="#B45-gels-10-00754" class="html-bibr">45</a>,<a href="#B46-gels-10-00754" class="html-bibr">46</a>].</p>
Full article ">Figure 7
<p>Examples of bilayer constructs utilizing poly(N-isopropylacrylamide) hyaluronan derivative (<b>left</b>, (<b>a</b>–<b>d</b>) [<a href="#B63-gels-10-00754" class="html-bibr">63</a>] and cellulose (<b>right</b>, (<b>A</b>–<b>C</b>) uncoated meshes/(<b>D</b>–<b>F</b>) bacterial cellulose-coated meshes) [<a href="#B84-gels-10-00754" class="html-bibr">84</a>].</p>
Full article ">Figure 8
<p>Evaluation methods for mechanical biocompatibility of surgical meshes: (<b>a</b>) uniaxial tensile, (<b>b</b>) biaxial tensile test, (<b>c</b>,<b>d</b>) ball bursting and deformation of mesh in ball burst testing, (<b>e</b>) suture retention, and (<b>f</b>) tear test [<a href="#B83-gels-10-00754" class="html-bibr">83</a>].</p>
Full article ">Figure 9
<p>Chemically (GDA) cross-linked (<b>left</b>) and physically (freeze-thawed) (<b>right</b>) PVA hydrogel implantation in Wistar rats (<b>a</b>,<b>b</b>) and Swine (<b>c</b>,<b>d</b>) models. Note, surgery on swine models was performed laparoscopically. Adapted from Dorkhani et al. and Fehér et al. [<a href="#B35-gels-10-00754" class="html-bibr">35</a>,<a href="#B91-gels-10-00754" class="html-bibr">91</a>].</p>
Full article ">
Previous Issue
Back to TopTop