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

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (7,385)

Search Parameters:
Keywords = hydrogels

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 2643 KiB  
Article
Enhancing the Efficacy of Radiation Therapy by Photochemical Internalization of Fibrin-Hydrogel-Delivered Bleomycin
by Sophia Renee Laurel, Keya Gupta, Jane Nguyen, Akhil Chandekar, Justin Le, Kristian Berg and Henry Hirschberg
Cancers 2024, 16(23), 4029; https://doi.org/10.3390/cancers16234029 (registering DOI) - 30 Nov 2024
Abstract
Background/Objectives: Although the use of radiation-sensitizing agents has been shown to enhance the effect of radiation on tumor cells, the blood–brain barrier (BBB) impedes these agents from reaching brain tumor sites when provided systemically. Localized methods of sensitizer delivery, utilizing hydrogels, have the [...] Read more.
Background/Objectives: Although the use of radiation-sensitizing agents has been shown to enhance the effect of radiation on tumor cells, the blood–brain barrier (BBB) impedes these agents from reaching brain tumor sites when provided systemically. Localized methods of sensitizer delivery, utilizing hydrogels, have the potential to bypass the blood–brain barrier. This study examined the ability of photochemical internalization (PCI) of hydrogel-released bleomycin to enhance the growth-inhibiting effects of radiation on multi-cell glioma spheroids in vitro. Methods: Loaded fibrin hydrogel layers were created by combining thrombin, fibrinogen, and bleomycin (BLM). Supernatants from these layers were collected, combined with photosensitizer, and added to F98 glioma spheroid cultures. Following light (PCI) and radiation treatment, at increasing dosages, spheroid growth was monitored for 14 days. Results: PCI of released BLM significantly reduced the radiation dose required to achieve equivalent efficacy compared to radiation or BLM + RT alone. Both immediate and delayed RT delivery post-BLM-PCI resulted in similar degrees of growth inhibition. Conclusions: Non-degraded BLM was released from the fibrin hydrogel. PCI of BLM synergistically increased the growth-inhibiting effects of radiation treatment compared to radiation and BLM, as well as radiation acting as a single treatment. Full article
(This article belongs to the Special Issue Novel Targeted Therapies in Brain Tumors)
16 pages, 3822 KiB  
Article
Detecting Hypoxia Through the Non-Invasive and Simultaneous Monitoring of Sweat Lactate and Tissue Oxygenation
by Cindy Cheng, Sayan Ganguly, Pei Li and Xiaowu Tang
Biosensors 2024, 14(12), 584; https://doi.org/10.3390/bios14120584 (registering DOI) - 30 Nov 2024
Viewed by 77
Abstract
Hypoxia, characterized by inadequate tissue oxygenation, may result in tissue damage and organ failure if not addressed. Current detection approaches frequently prove insufficient, depending on symptoms and rudimentary metrics such as tissue oxygenation, which fail to comprehensively identify the onset of hypoxia. The [...] Read more.
Hypoxia, characterized by inadequate tissue oxygenation, may result in tissue damage and organ failure if not addressed. Current detection approaches frequently prove insufficient, depending on symptoms and rudimentary metrics such as tissue oxygenation, which fail to comprehensively identify the onset of hypoxia. The European Pressure Ulcer Advisory Panel (EPUAP) has recognized sweat lactate as a possible marker for the early identification of decubitus ulcers, nevertheless, neither sweat lactate nor oxygenation independently provides an appropriate diagnosis of hypoxia. We have fabricated a wearable device that non-invasively and concurrently monitors sweat lactate and tissue oxygenation to fill this gap. The apparatus comprises three essential components: (i) a hydrogel-based colorimetric lactate biosensor, (ii) a near-infrared (NIR) sensor for assessing tissue oxygenation, and (iii) an integrated form factor for enhanced wearability. The lactate sensor alters its hue upon interaction with lactate in sweat, whereas the NIR sensor monitors tissue oxygenation levels in real-time. The device underwent testing on phantom exhibiting tissue-mimicking characteristics and on human sweat post aerobic and anaerobic activities. Moreover, the device was demonstrated to be capable of real-time “on-body” simultaneous monitoring of sweat lactate spikes and tissue oxygenation (StO2) drops, which showed strong correlation during a hypoxia protocol. This innovative technology has a wide range of potential applications, such as post-operative care, sepsis detection, and athletic performance monitoring, and may provide economical healthcare solutions in resource-limited regions. Full article
(This article belongs to the Special Issue Biosensors for Monitoring and Diagnostics)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of (<b>a</b>) the lactate-sensing hydrogel composition and (<b>b</b>) the colorimetric sensing mechanism. Oxidation of lactate, catalyzed by LOX, generates H<sub>2</sub>O<sub>2</sub> as a byproduct; TMB reacts with H<sub>2</sub>O<sub>2</sub>, catalyzed by HRP, results in a color change. (<b>c</b>) Optical absorption peaks of TMB (350 nm) and TMB<sub>ox</sub> (one-electron oxidation, 660 and 960 nm; two-electron oxidation, 450 nm).</p>
Full article ">Figure 2
<p>(<b>a</b>) Images of gels in a multi-well plate in response to various lactate concentrations at various time points. (<b>b</b>) Representative optical absorption spectra of gels in response to lactate.</p>
Full article ">Figure 3
<p>(<b>a</b>) An exploded view of the device prototype. Pictures of an assembled wearable device, (<b>b</b>) without, and (<b>c</b>) with the gel. Optical pathways from LEDs to PDs (<b>d</b>) without and (<b>e</b>) with the lactate sensing hydrogel on tissue. Left: simulation in tissue; right: bottom view of the wearable device.</p>
Full article ">Figure 4
<p>(<b>a</b>) Picture of the device placed on top of an optical phantom with optical properties mimicking those of human tissue (StO<sub>2</sub> = 68%). (<b>b</b>) Normalized absorption of the lactate sensing hydrogels, extracted from the optical readout, in response to various lactate solutions (0, 10, 40, 80 mM) in phosphate buffer (pH 6.0) respectively. (<b>c</b>,<b>d</b>) Stable StO<sub>2</sub> and BVI reading regardless of lactate concentration.</p>
Full article ">Figure 5
<p>(<b>a</b>) Picture of the test setup showing the device worn on the researcher’s left thigh while on a stational bike, with the laptop running data collection software. (<b>b</b>) Normalized absorption vs. time response curves for the aerobic and anaerobic conditions. Inset-left up: appearance of gel before and after the test; inset-right down: lactate detection calibration curve.</p>
Full article ">Figure 6
<p>(<b>a</b>) Picture of a student researcher sitting in a sauna tent to stimulate sweating while a pressure cuff was used to create temporary hypoxia in the forearm. (<b>b</b>) A zoomed-in view of the pressure cuff worn on the upper arm and the wearable device on the forearm. (<b>c</b>–<b>e</b>) Simultaneous monitoring of sweat lactate production (top panel), tissue oxygenation level (middle panel), and BVI (bottom panel).</p>
Full article ">
23 pages, 8866 KiB  
Article
New Carrageenan/2-Dimethyl Aminoethyl Methacrylate/Gelatin/ZnO Nanocomposite as a Localized Drug Delivery System with Synergistic Biomedical Applications
by Abeer A. Ageeli and Sahera F. Mohamed
Processes 2024, 12(12), 2702; https://doi.org/10.3390/pr12122702 (registering DOI) - 30 Nov 2024
Viewed by 387
Abstract
In recent years, the development of multifunctional hydrogels has gained significant attention due to their potential in various biomedical applications, including antimicrobial, antioxidant, and anticancer therapies. By integrating biocompatible polymers and nanoparticles, these hydrogels can achieve enhanced activity and targeted therapeutic effects. In [...] Read more.
In recent years, the development of multifunctional hydrogels has gained significant attention due to their potential in various biomedical applications, including antimicrobial, antioxidant, and anticancer therapies. By integrating biocompatible polymers and nanoparticles, these hydrogels can achieve enhanced activity and targeted therapeutic effects. In this study, carrageenan/2-dimethyl aminoethyl methacrylate/gelatin (CAR/DEMA/Gelt) composite hydrogel was synthesized using microwave radiation specifically for its efficiency in enhancing cross-linking and promoting uniform nanoparticle dispersion within the matrix. Zinc oxide (ZnO) nanoparticles were incorporated into the hydrogel to form the (CAR/DEMA/Gelt/ZnO) nanocomposite. The hydrogels were characterized using FT-IR, FE-SEM, XRD, TGA, and EDX, confirming successful cross-linking and structural integrity. The nanocomposite hydrogel exhibited more enhanced antimicrobial activity than the composite hydrogel against Gram-positive Staphylococcus aureus (S. aureus) and Bacillus subtilis (B. subtilis), with inhibition zones of 15 mm and 16 mm, respectively, while in case of the Gram-negative bacteria, Klebsiella pneumoniae (K. pneumoniae) and Escherichia coli (E. coli), the inhibition zones were 29 mm and 19 mm, respectively. In addition to the unicellular fungi, Candida albicans (C. albicans), the inhibition zone was 19 mm. Moreover, the nanocomposite showed anti-inflammatory activity comparable to those of Indomethacin and antioxidant activity, with an impressive IC50 value of 33.3 ± 0.05 µg/mL. In vitro cytotoxicity assays revealed significant anticancer activity. Against the MCF-7 breast cancer cell line, the CAR/DEMA/Gelt/ZnO nanocomposite showed 72.5 ± 0.02% cell viability, which decreased to 30.8 ± 0.01% after loading doxorubicin (DOX). Similarly, against the HepG2 liver cancer cell line, the free nanocomposite displayed 59.9 ± 0.006% cell viability, which depleted to 29.9 ± 0.005% when DOX was uploaded. This CAR/DEMA/Gelt/ZnO nanocomposite hydrogel demonstrates strong potential as a multifunctional platform for targeted biomedical applications, particularly in cancer therapy. Full article
Show Figures

Figure 1

Figure 1
<p>FTIR analysis of (<b>a</b>) ZnO nanoparticles, (<b>b</b>) DOX powder, and (<b>c</b>) CAR/DEMA/Gelt hydrogel, CAR/DEMA/Gelt/ZnO nanocomposite, and CAR/DEMA/Gelt/ZnO/DOX.</p>
Full article ">Figure 1 Cont.
<p>FTIR analysis of (<b>a</b>) ZnO nanoparticles, (<b>b</b>) DOX powder, and (<b>c</b>) CAR/DEMA/Gelt hydrogel, CAR/DEMA/Gelt/ZnO nanocomposite, and CAR/DEMA/Gelt/ZnO/DOX.</p>
Full article ">Figure 2
<p>XRD patterns of (<b>a</b>) ZnO nanoparticles and (<b>b</b>) CAR/DEMA/Gelt hydrogel and CAR/DEMA/Gelt/ZnO nanocomposite.</p>
Full article ">Figure 3
<p>TGA (<b>a</b>) and DTA (<b>b</b>) analysis of CAR/DEMA/Gelt hydrogel and CAR/DEMA/Gelt/ZnO nanocomposite.</p>
Full article ">Figure 4
<p>FE-SEM of CAR/DEMA/Gelt hydrogel, CAR/DEMA/Gelt/ZnO nanocomposite, and CAR/DEMA/Gelt/ZnO/DOX nanocomposite.</p>
Full article ">Figure 5
<p>(<b>a</b>) EDX elemental analysis and mapping of the CAR/DEMA/Gelt hydrogel. (<b>b</b>) EDX elemental analysis and mapping of the CAR/DEMA/Gelt/ZnO nanocomposite hydrogel.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) EDX elemental analysis and mapping of the CAR/DEMA/Gelt hydrogel. (<b>b</b>) EDX elemental analysis and mapping of the CAR/DEMA/Gelt/ZnO nanocomposite hydrogel.</p>
Full article ">Figure 6
<p>Antimicrobial activity of the CAR/DEMA/Gelt hydrogel (i) and the CAR/DEMA/Gelt/ZnO nanocomposite (ii).</p>
Full article ">Figure 6 Cont.
<p>Antimicrobial activity of the CAR/DEMA/Gelt hydrogel (i) and the CAR/DEMA/Gelt/ZnO nanocomposite (ii).</p>
Full article ">Figure 7
<p>Anti-inflammatory activity of the CAR/DEMA/Gelt/ZnO nanocomposite using Indomethacin as a control.</p>
Full article ">Figure 8
<p>Antioxidant activity of the CAR/DEMA/Gelt/ZnO nanocomposite using ascorbic acid as a control.</p>
Full article ">Figure 9
<p>(<b>a</b>) The cell viability and cytotoxicity (%) data for CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX against MCF-7. (<b>b</b>) Effects of CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX on MCF-7 cells.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) The cell viability and cytotoxicity (%) data for CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX against MCF-7. (<b>b</b>) Effects of CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX on MCF-7 cells.</p>
Full article ">Figure 10
<p>(<b>a</b>) The cell viability and cytotoxicity (%) data for CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX against the HepG2 cell line. (<b>b</b>) Effects of CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX on HepG2 cells.</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>) The cell viability and cytotoxicity (%) data for CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX against the HepG2 cell line. (<b>b</b>) Effects of CAR/DEMA/Gelt/ZnO and CAR/DEMA/Gelt/ZnO/DOX on HepG2 cells.</p>
Full article ">
19 pages, 4159 KiB  
Article
Microwave-Assisted Production of Defibrillated Lignocelluloses from Blackcurrant Pomace via Citric Acid and Acid-Free Conditions
by Natthamon Inthalaeng, Ryan E. Barker, Tom I. J. Dugmore and Avtar S. Matharu
Molecules 2024, 29(23), 5665; https://doi.org/10.3390/molecules29235665 (registering DOI) - 29 Nov 2024
Viewed by 167
Abstract
Blackcurrant pomace (BCP) is an example of an annual, high-volume, under-utilized renewable resource with potential to generate chemicals, materials and bioenergy within the context of a zero-waste biorefinery. Herein, the microwave-assisted isolation, characterization and potential application of defibrillated lignocelluloses from depectinated blackcurrant pomace [...] Read more.
Blackcurrant pomace (BCP) is an example of an annual, high-volume, under-utilized renewable resource with potential to generate chemicals, materials and bioenergy within the context of a zero-waste biorefinery. Herein, the microwave-assisted isolation, characterization and potential application of defibrillated lignocelluloses from depectinated blackcurrant pomace are reported. Depectination was achieved using citric acid (0.2–0.8 M, 80 °C, 2 h, conventional heating) and compared with acid-free hydrothermal microwave-assisted processing (1500 W, 100–160 °C, 30 min). The resultant depectinated residues were subjected to microwave-assisted hydrothermal defibrillation to afford two classes of materials: namely, (i) hydrothermal acid-free microwave-assisted (1500 W, 160 °C, 30 min; DFC-M1-M4), and (ii) hydrothermal citric acid microwave-assisted (1500 W, 160 °C, 30 min; DFC-C1–C4). Thermogravimetric analysis (TGA) revealed that the thermal stability with respect to native BCP (Td = 330 °C) was higher for DFC-M1-M4 (Td = 345–348 °C) and lower for DFC-C1–C4 (322–325 °C). Both classes of material showed good propensity to hold water but failed to form stable hydrogels (5–7.5 wt% in water) unless they underwent bleaching which removed residual lignin and hemicellulosic matter, as evidenced by 13C solid-state NMR spectroscopy. The hydrogels made from bleached DFC-C1–C4 (7.5 wt%) and bleached DFC-M1-M4 (5 wt%) exhibited rheological viscoelastic, shear thinning, and time-dependent behaviour, which highlights the potential opportunity afforded by microwave-assisted defibrillation of BCP for food applications. Full article
14 pages, 3383 KiB  
Article
Fabrication and Characterization of Phyllanthus Emblica Extract-Polyvinyl Alcohol/Carboxymethyl Cellulose Sodium Antioxidant Hydrogel and Its Application in Wound Healing
by Shanqin Huang, Shanglun Li, Guoyan Li, Chenyu Wang, Xiaohan Guo, Jing Zhang, Jing Liu, Ying Xu and Yanchun Wang
Pharmaceutics 2024, 16(12), 1531; https://doi.org/10.3390/pharmaceutics16121531 - 29 Nov 2024
Viewed by 178
Abstract
Background: Phyllanthus emblica is a medicinal and edible plant from the Euphorbiaceae family, notable for its rich content of polyphenols and flavonoids, which provide significant antioxidant properties. To exploit the full antioxidant potential of Phyllanthus emblica, this study developed a hydrogel system [...] Read more.
Background: Phyllanthus emblica is a medicinal and edible plant from the Euphorbiaceae family, notable for its rich content of polyphenols and flavonoids, which provide significant antioxidant properties. To exploit the full antioxidant potential of Phyllanthus emblica, this study developed a hydrogel system incorporating polyvinyl alcohol (PVA) and carboxymethyl cellulose sodium (CMC-Na), integrated with Phyllanthus emblica extract, for the purpose of wound healing. Methods: The extraction process of active ingredients of Phyllanthus emblica was optimized and assessed the antioxidant composition and activity of the extract. A series of hydrogel performance evaluations were performed on the Phyllanthus emblica extract-loaded PVA/CMC-Na hydrogel (AEPE composite hydrogel). Additionally, the wound healing efficacy was evaluated through cell culture experiments and wound healing assays using BALB/C mice. Results: The findings indicated that the extraction of Phyllanthus emblica with 95% ethanol yielded an extract rich in polyphenols, primarily gallic acid and ellagic acid, demonstrating high free radical scavenging capacity and robust antioxidant activity. The hydrogel matrix containing 12% PVA and 1% CMC-Na exhibited excellent physicochemical properties. The optimized AEPE composite hydrogel enabled sustained drug release over a 24 h period, exhibited low cytotoxicity and promoted cell migration. In a mouse dorsal wound healing model, the AEPE composite hydrogel showed pronounced anti-inflammatory and antioxidation effects, enhanced collagen deposition, and ultimately accelerated wound healing. Conclusions: The AEPE composite hydrogel demonstrated strong antioxidant characteristics and significant wound healing potential. Thus, this study could broaden the application prospects of Phyllanthus emblica in wound healing. Full article
(This article belongs to the Section Physical Pharmacy and Formulation)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Contents of total polyphenols in water extract (WEPE) and ethanol extract (AEPE) of Phyllanthus emblica; (<b>b</b>) scavenging rate of DPPH, ABTS<sup>+</sup>, and hydroxyl radical by Phyllanthus emblica officinalis extracts; HPLC of gallic acid (<b>c</b>), ellagic acid (<b>d</b>), and AEPE (<b>e</b>). # represents the difference between each solvent extraction and gallic acid (# <span class="html-italic">p</span> &lt; 0.05, ### <span class="html-italic">p</span> &lt; 0.001); * represents the difference between individual solvent extracts (mean ± SD, <span class="html-italic">n</span> = 3; * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 2
<p>Evaluation of AEPE composite hydrogel performance: (<b>a</b>) residual weight rate of hydrogel with different ratios at 24 h and (<b>b</b>) 72 h; (<b>c</b>) maximum swelling rate of hydrogel with different ratios; (<b>d</b>) water retention of hydrogel with different ratios; (<b>e</b>) wetting rate of hydrogel with different ratios; (<b>f</b>) hydrogel fraction of hydrogel with different ratios. # represents the difference between CMC-Na with each content and 0% CMC-Na (# <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); * represents the difference between the CMC-Na of each content (mean ± SD, <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).</p>
Full article ">Figure 3
<p>Scanning electron microscopy images of (<b>a</b>) PVA/CMC-Na hydrogel and (<b>b</b>) AEPE composite hydrogel (** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001); (<b>c</b>) free radical scavenging of the AEPE composite hydrogel; (<b>d</b>) in vitro release curves of PVA/CMC-Na hydrogel with different gallic acid contents over 24 h. Rheological tests: (<b>e</b>) shear viscosity and (<b>f</b>) frequency scanning.</p>
Full article ">Figure 4
<p>(<b>a</b>) Cell viabilities of 3T6-Swiss albino mice treated with free film for 24 h and 48 h; (<b>b</b>) rates and (<b>c</b>) fibroblast migration behaviors after blank hydrogel with AEPE hydrogel. # represents the difference between hydrogel of each group and control group (### <span class="html-italic">p</span> &lt; 0.001); * represents the difference between the hydrogels of each group (mean ± SD, <span class="html-italic">n</span> = 3; ** <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>) The wound healing processes recorded for the rats by using different treatments (control, AEPE, blank hydrogel, and AEPE hydrogel) at various times (0–14 days); (<b>b</b>) time-dependent wound healing area after different treatments. (<b>c</b>) Histological analysis of tumor sections on day 14 after different treatments. (<b>d</b>) TNF-α, IL-1β, and IL-6 concentrations, expressed as pg/mL, in the control and treatment groups (d7). # represents the difference between the treatment groups and the control group (### <span class="html-italic">p</span> &lt; 0.001); * represents the difference between treatment groups (mean ± SD, <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).</p>
Full article ">
18 pages, 5815 KiB  
Article
Aluminum-Free Borosilicate Glass Functionalized Hydrogels for Enhanced Dental Tissue Regeneration
by Nina Attik, Inès Basri, Jérôme Sohier, Rémy Gauthier, Cyril Villat and Christelle Goutaudier
Materials 2024, 17(23), 5862; https://doi.org/10.3390/ma17235862 - 29 Nov 2024
Viewed by 233
Abstract
Hydrogels are promising scaffolds for tissue regeneration, and borosilicate glass particles have demonstrated potential in enhancing the biological behaviour of dental pulp cells. However, the specific morphological characteristics of dental lesions and the diverse requirements of dental tissues require biocompatible, bioactive, and shapeable [...] Read more.
Hydrogels are promising scaffolds for tissue regeneration, and borosilicate glass particles have demonstrated potential in enhancing the biological behaviour of dental pulp cells. However, the specific morphological characteristics of dental lesions and the diverse requirements of dental tissues require biocompatible, bioactive, and shapeable scaffolds. This study aimed to evaluate the in vitro biological behaviour of human gingival fibroblasts (HGFs) in contact with an experimental aluminum-free borosilicate glass-functionalized hydrogel. Two types of experimental borosilicate glass particles were utilized, with Biodentine® particles serving as a reference material. The hydrogel, based on poly(L-lysine) dendrimers (DGL) with or without borosilicate particles, was analyzed using micro-computed tomography (µCT) and scanning electron microscopy (SEM) coupled with energy-dispersive X-ray spectroscopy (EDX). Cytocompatibility was assessed using Live/Dead™ staining, and cell colonization was evaluated via confocal imaging. Additionally, Alizarin red staining was performed to assess mineralization potential after 7 and 14 days. Results indicated that the incorporation of borosilicate particles did not alter hydrogel porosity, while EDX confirmed particle presence on the hydrogel surfaces. Furthermore, the borosilicate-functionalized hydrogels significantly enhanced cell proliferation, colonization, and the content of calcium deposits. These findings highlight the potential of these hydrogels for future clinical applications in dental tissue regeneration, pending further development. Full article
Show Figures

Figure 1

Figure 1
<p>SEM image of functionalized hydrogels compared to the control hydrogel group (scale bars = 100 µm). Yellow circles indicate the presence of particles on the hydrogel surfaces.</p>
Full article ">Figure 2
<p>EDX mapping images of the tested functionalized gels: (<b>1</b>) Control gel, (<b>2</b>) Biodentine<sup>®</sup> gel, (<b>3</b>) PSBS3 gel, and (<b>4</b>) PSBS8 gel, illustrating the morphology of various elements and their distribution profiles. The corresponding mapping areas are shown on the left side.</p>
Full article ">Figure 3
<p>pH variations in control and functionalized hydrogels over time.</p>
Full article ">Figure 4
<p>Time-dependent changes in calcium concentration. Data are mean ± SD of 3 independent experiments (n = 9). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001 denote significant differences.</p>
Full article ">Figure 5
<p>3D reconstructions of the tested hydrogels using ARIVIS 4.3 software: (<b>A</b>) control hydrogel, (<b>B</b>) Biodentine<sup>®</sup> hydrogel, (<b>C</b>) PSBS8 hydrogel, and (<b>D</b>) PSBS3 hydrogel (scale bares = 100 µm).</p>
Full article ">Figure 6
<p>Representative TGA curve of the tested gels.</p>
Full article ">Figure 7
<p>Representative DSC curve of the tested gels. (<b>A</b>) Peak temperature of crystallization, (<b>B</b>) glass transition temperature, and (<b>C</b>) peak temperature of melting.</p>
Full article ">Figure 8
<p>Assessment of HGF cytotoxicity using Live/Dead staining. Cells were observed under an epifluorescence microscope after 24 h of direct contact. Green areas indicate live cells, while red areas indicate damaged cells. No damaged (red) cells were observed. (Scale bars = 20 µm).</p>
Full article ">Figure 9
<p>Spreading and colonization of hydrogels by HGF cells observed via confocal laser imaging after 24 h of direct contact with the various hydrogels (combination of DAPI-nucleus channels in Blue and Alexa Fluor<sup>TM</sup> 488 Phalloidin. Cytoskeleton in green. Scale bars = 50 µm).</p>
Full article ">Figure 10
<p>Quantification of calcium deposition by Alizarin red staining in HGF cells (B) as a function of contact time and hydrogel type. Data represent mean ± SD of 3 independent experiments (n = 6). * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001 indicate statistically significant values.</p>
Full article ">
50 pages, 19208 KiB  
Review
Nanotechnology in Drug Delivery: Anatomy and Molecular Insight into the Self-Assembly of Peptide-Based Hydrogels
by Adelaide R. Mashweu and Vladimir A. Azov
Molecules 2024, 29(23), 5654; https://doi.org/10.3390/molecules29235654 - 29 Nov 2024
Viewed by 358
Abstract
The bioavailability, release, and stability of pharmaceuticals under physicochemical conditions is the major cause of drug candidates failing during their clinical trials. Therefore, extensive efforts have been invested in the development of novel drug delivery systems that are able to transport drugs to [...] Read more.
The bioavailability, release, and stability of pharmaceuticals under physicochemical conditions is the major cause of drug candidates failing during their clinical trials. Therefore, extensive efforts have been invested in the development of novel drug delivery systems that are able to transport drugs to a desired site and improve bioavailability. Hydrogels, and peptide hydrogels in particular, have been extensively investigated due to their excellent biocompatibility and biodegradability properties. However, peptide hydrogels often have weak mechanical strength, which limits their therapeutic efficacy. Therefore, a number of methods for improving their rheological properties have been established. This review will cover the broad area of drug delivery, focusing on the recent developments in this research field. We will discuss the variety of different types of nanocarrier drug delivery systems and then, more specifically, the significance and perspectives of peptide-based hydrogels. In particular, the interplay of intermolecular forces that govern the self-assembly of peptide hydrogels, progress made in understanding the distinct morphologies of hydrogels, and applications of non-canonical amino acids in hydrogel design will be discussed in more detail. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Drug absorption through oral admission. Reprinted with permission from Ref. [<a href="#B10-molecules-29-05654" class="html-bibr">10</a>]. Copyright 2023 MDPI.</p>
Full article ">Figure 2
<p>Skin structure depicting different skin layers. Reprinted with permission from Ref. [<a href="#B16-molecules-29-05654" class="html-bibr">16</a>]. Copyright 2015 MDPI.</p>
Full article ">Figure 3
<p>Limitations of conventional drug delivery systems.</p>
Full article ">Figure 4
<p>Examples of coumarin derivatives <b>1a</b>–<b>8a</b> [<a href="#B28-molecules-29-05654" class="html-bibr">28</a>].</p>
Full article ">Figure 5
<p>Activation of prodrugs by a cleavage reaction. Reprinted with permission from Ref. [<a href="#B36-molecules-29-05654" class="html-bibr">36</a>]. Copyright 2020 MDPI.</p>
Full article ">Figure 6
<p>Various structures of nanosized materials. Reprinted with permission from Ref. [<a href="#B56-molecules-29-05654" class="html-bibr">56</a>]. Copyright 2015 American Chemical Society.</p>
Full article ">Figure 7
<p>Different categories of nanocarriers.</p>
Full article ">Figure 8
<p>Organic nanocarriers [<a href="#B67-molecules-29-05654" class="html-bibr">67</a>].</p>
Full article ">Figure 9
<p>Doxorubicin (DOX), commonly used as an anti-cancer drug.</p>
Full article ">Figure 10
<p>Passive and targeted drug delivery of nanocarriers. The targeted drug delivery nanocarrier has residues complementary to the binding sites on the targeted cell surface. Reprinted with permission from Ref. [<a href="#B84-molecules-29-05654" class="html-bibr">84</a>]. Copyright 2018 American Chemical Society.</p>
Full article ">Figure 11
<p>Factors contributing to the release of drugs from nanocarriers. Reprinted with permission from Ref. [<a href="#B85-molecules-29-05654" class="html-bibr">85</a>]. Copyright 2015 Elsevier.</p>
Full article ">Figure 12
<p>N-(1,3-dihydroxypropan-2-yl) methacrylamide (DHPMA) conjugated with DOX. (<b>A</b>) Molecular structure; (<b>B</b>) nanocarrier self-assembly; (<b>C</b>) cellular function. Reprinted with permission from Ref. [<a href="#B77-molecules-29-05654" class="html-bibr">77</a>]. Copyright 2019 Elsevier.</p>
Full article ">Figure 13
<p>Three-dimensional network in hydrogels and their applications in medicine. Reprinted with permission from Ref. [<a href="#B92-molecules-29-05654" class="html-bibr">92</a>]. Copyright 2021 Frontiers.</p>
Full article ">Figure 14
<p>Cross-linking of polymers in gels. Reprinted with permission from Ref. [<a href="#B109-molecules-29-05654" class="html-bibr">109</a>]. Copyright 2021 MDPI.</p>
Full article ">Figure 15
<p>Synthetic polymers used in the preparation of polymeric hydrogels [<a href="#B130-molecules-29-05654" class="html-bibr">130</a>].</p>
Full article ">Figure 16
<p>Drugs physically and chemically loaded within a hydrogel. Reprinted with permission from Ref. [<a href="#B13-molecules-29-05654" class="html-bibr">13</a>]. Copyright 2019 Springer Nature.</p>
Full article ">Figure 17
<p>High affinity of MG-1 towards CIP·HCl [<a href="#B137-molecules-29-05654" class="html-bibr">137</a>].</p>
Full article ">Figure 18
<p>Cross-linking of chemical hydrogels via dynamic covalent bonds. Reprinted with permission from Refs. [<a href="#B97-molecules-29-05654" class="html-bibr">97</a>,<a href="#B142-molecules-29-05654" class="html-bibr">142</a>]. Copyright 2020 and 2022 MDPI.</p>
Full article ">Figure 19
<p>In situ gelation of physical and chemical hydrogels. Reprinted with permission from Ref. [<a href="#B89-molecules-29-05654" class="html-bibr">89</a>]. Copyright 2018 Springer Nature.</p>
Full article ">Figure 20
<p>Different release modes of loaded drug from hydrogel. Reprinted with permission from Refs. [<a href="#B91-molecules-29-05654" class="html-bibr">91</a>,<a href="#B95-molecules-29-05654" class="html-bibr">95</a>]. Copyright 2021 Wiley, 2008 Elsevier.</p>
Full article ">Figure 21
<p>Application of peptide hydrogels. Reprinted with permission from Ref. [<a href="#B135-molecules-29-05654" class="html-bibr">135</a>]. Copyright 2021 Frontiers.</p>
Full article ">Figure 22
<p>Structures of 20 canonical amino acids.</p>
Full article ">Figure 23
<p>L and D enantiomers of amino acids.</p>
Full article ">Figure 24
<p>α-helix and β-sheet secondary structures of peptides [<a href="#B175-molecules-29-05654" class="html-bibr">175</a>]. Image credit: OpenStax Biology.</p>
Full article ">Figure 25
<p>Classical examples of amphipathic peptide hydrogelators [<a href="#B96-molecules-29-05654" class="html-bibr">96</a>,<a href="#B190-molecules-29-05654" class="html-bibr">190</a>].</p>
Full article ">Figure 26
<p>The self-assembly of peptides leads to hydrogelation. Reprinted with permission from Ref. [<a href="#B102-molecules-29-05654" class="html-bibr">102</a>]. Copyright 2013 Royal Society of Chemistry.</p>
Full article ">Figure 27
<p>Structure of H-FQFQFK-NH<sub>2</sub> and its dimer [<a href="#B24-molecules-29-05654" class="html-bibr">24</a>].</p>
Full article ">Figure 28
<p>Influence of ionic strength on self-assembly of K2 peptide. Reprinted with permission from Ref. [<a href="#B194-molecules-29-05654" class="html-bibr">194</a>]. Copyright 2024 American Chemical Society.</p>
Full article ">Figure 29
<p>Histidine-based peptide amphiphiles, which disassemble under acidic environments [<a href="#B195-molecules-29-05654" class="html-bibr">195</a>].</p>
Full article ">Figure 30
<p>Halogenated phenyl groups (X = halogen) in a peptide sequence to influence the strength of peptide hydrogels [<a href="#B200-molecules-29-05654" class="html-bibr">200</a>].</p>
Full article ">Figure 31
<p>Self-assembly kinetic mechanisms proposed by MD simulation of amphiphilic peptide C<sub>16</sub>H<sub>31</sub>O-(VaI)<sub>3</sub>-(AIa)<sub>3</sub>-(GIu)<sub>3</sub> in aqueous solution (<b>A</b>) at weak hydrophobic interaction strength networks of β-sheets are formed; (<b>B</b>) at moderate hydrophobic interaction strength cylindrical nanofibers are formed; (<b>C</b>) at strong hydrophobic interaction strength elongated micelles are formed. Reprinted with permission from Ref. [<a href="#B202-molecules-29-05654" class="html-bibr">202</a>]. Copyright 2015 Langmuir.</p>
Full article ">Figure 32
<p>A family of dibenzoyl-L-cystine (<b>1h</b>) derivatives demonstrating hydrogelation properties [<a href="#B203-molecules-29-05654" class="html-bibr">203</a>].</p>
Full article ">Figure 33
<p>Different phases during fibril formation as progressed over time from the monomer phase to fully assembled phase. Reprinted with permission from Ref. [<a href="#B211-molecules-29-05654" class="html-bibr">211</a>]. Copyright 2017 Elsevier.</p>
Full article ">Figure 34
<p>Representative examples of non-canonical amino acids [<a href="#B162-molecules-29-05654" class="html-bibr">162</a>].</p>
Full article ">Figure 35
<p>Halogenated Fmoc-Phe hydrogelators [<a href="#B223-molecules-29-05654" class="html-bibr">223</a>].</p>
Full article ">Figure 36
<p>Fmoc-Phe derivatives [<a href="#B183-molecules-29-05654" class="html-bibr">183</a>].</p>
Full article ">Figure 37
<p>L-azetidine-2-carboxilic acid found in sugar beets [<a href="#B232-molecules-29-05654" class="html-bibr">232</a>].</p>
Full article ">Scheme 1
<p>Generation of the Acyclovir active drug in vivo [<a href="#B41-molecules-29-05654" class="html-bibr">41</a>].</p>
Full article ">Scheme 2
<p>Levodopa as a bioprecursor prodrug for dopamine [<a href="#B41-molecules-29-05654" class="html-bibr">41</a>].</p>
Full article ">Scheme 3
<p>Carrier phosphate prodrugs for atazanavir [<a href="#B39-molecules-29-05654" class="html-bibr">39</a>].</p>
Full article ">Scheme 4
<p>Two activation steps for the release of <b>2e</b> [<a href="#B39-molecules-29-05654" class="html-bibr">39</a>].</p>
Full article ">Scheme 5
<p>Metabolic pathways of codeine [<a href="#B43-molecules-29-05654" class="html-bibr">43</a>].</p>
Full article ">Scheme 6
<p>Disassembly and assembly of H-bonding interactions in physical hydrogels are shown as an example of binding between methylcellulose and hyaluronic acid. Reprinted with permission from Ref. [<a href="#B95-molecules-29-05654" class="html-bibr">95</a>]. Copyright 2008 Elsevier.</p>
Full article ">Scheme 7
<p>SPPS based on Boc and Fmoc chemistry.</p>
Full article ">Scheme 8
<p>Three-dimensional network formation of peptide hydrogels. Reprinted with permission from Refs. [<a href="#B24-molecules-29-05654" class="html-bibr">24</a>,<a href="#B53-molecules-29-05654" class="html-bibr">53</a>,<a href="#B184-molecules-29-05654" class="html-bibr">184</a>]. Copyright 2020 MDPI, 2012 and 2022 Elsevier.</p>
Full article ">Scheme 9
<p>Synthetic methodologies for ncAAs [<a href="#B192-molecules-29-05654" class="html-bibr">192</a>,<a href="#B229-molecules-29-05654" class="html-bibr">229</a>].</p>
Full article ">Scheme 10
<p>Possible approaches for ncAAs synthesis [<a href="#B227-molecules-29-05654" class="html-bibr">227</a>].</p>
Full article ">Scheme 11
<p>Negishi cross-coupling reaction with a protected iodoserine derivative [<a href="#B230-molecules-29-05654" class="html-bibr">230</a>].</p>
Full article ">
17 pages, 4747 KiB  
Article
Physicochemical Properties, Drug Release and In Situ Depot-Forming Behaviors of Alginate Hydrogel Containing Poorly Water-Soluble Aripiprazole
by Hy D. Nguyen, Munsik Jang, Hai V. Ngo, Myung-Chul Gil, Gang Jin, Jing-Hao Cui, Qing-Ri Cao and Beom-Jin Lee
Gels 2024, 10(12), 781; https://doi.org/10.3390/gels10120781 (registering DOI) - 29 Nov 2024
Viewed by 226
Abstract
The objective of this study was to investigate the physicochemical properties, drug release and in situ depot-forming behavior of alginate hydrogel containing poorly water-soluble aripiprazole (ARP) for achieving free-flowing injectability, clinically accessible gelation time and sustained drug release. The balanced ratio of pyridoxal [...] Read more.
The objective of this study was to investigate the physicochemical properties, drug release and in situ depot-forming behavior of alginate hydrogel containing poorly water-soluble aripiprazole (ARP) for achieving free-flowing injectability, clinically accessible gelation time and sustained drug release. The balanced ratio of pyridoxal phosphate (PLP) and glucono-delta-lactone (GDL) was crucial to modulate gelation time of the alginate solution in the presence of calcium carbonate. Our results demonstrated that the sol state alginate hydrogel before gelation was free-flowing, stable and readily injectable using a small 23 G needle. In addition, the ratio (w/w) of PLP and GDL altered the gelation time, which was longer as the PLP content increased but shorter as the GDL content increased. The alginate hydrogel with a ratio of PLP to GDL of 15:9 had the optimal physicochemical properties in terms of a clinically acceptable gelation time (9.1 min), in situ-depot formation with muscle-mimicking stiffness (3.55 kPa) and sustained release over a two-week period. The alginate hydrogel, which is tunable by varying the ratio of PLP and GDL, could provide a controllable pharmaceutical preparation to meet the need for long-acting performance of antipsychotic drugs like ARP. Full article
Show Figures

Figure 1

Figure 1
<p>Physical appearance of alginate hydrogel in (<b>a</b>) the sol state and (<b>b</b>) the gel state, and (<b>c</b>) the hydrogel containing a fixed ARP dose of 14 mg.</p>
Full article ">Figure 2
<p>Gelation time of alginate hydrogel formulations. Data are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3). The significant differences were analyzed compared with F1 by Student’s <span class="html-italic">t</span>-test: ** <span class="html-italic">p</span> &lt; 0.01, * <span class="html-italic">p</span> &lt; 0.05, and ns: no significant difference (<span class="html-italic">p</span> &gt; 0.05). A detailed statistical analysis is presented in <a href="#app1-gels-10-00781" class="html-app">Table S1</a>.</p>
Full article ">Figure 3
<p>SEM images of the surface morphologies (×50) of dried alginate hydrogels after the gelation process: (<b>a</b>) F1, (<b>b</b>) F3, (<b>c</b>) F7 and (<b>d</b>) F8. Cross-sectional morphologies (×200) of (<b>e</b>) F1, (<b>f</b>) F3, (<b>g</b>) F7 and (<b>h</b>) F8.</p>
Full article ">Figure 4
<p>Swelling ratio of alginate hydrogel formulations. Data are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3). The significant differences at 24 h were analyzed by Student’s <span class="html-italic">t</span>-test. ns: no significant difference (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 5
<p>Degradation rate of alginate hydrogel formulations. Data are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3). The significant differences were analyzed on Day 6 and Day 14 by Student’s <span class="html-italic">t</span>-test; ns: no significant difference (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>Compressive stress–strain curve of alginate hydrogel formulations. Data are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>In vitro dissolution profiles of Abilify Maintena<sup>®</sup>, pure ARP powder and alginate hydrogels using (<b>a</b>) USP Apparatus II and (<b>b</b>) the shaking bath method. Data are expressed as the mean ± standard deviation (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 8
<p>Viscosity of the pregel solution (F8) after 1 month of storing at 25 °C.</p>
Full article ">Figure 9
<p>(<b>a</b>) Injectability of the sol state using a 23 G needle and (<b>b</b>,<b>c</b>) in situ depot-forming behavior of alginate hydrogel (F8) in pig muscle, showing (<b>b</b>) surface and (<b>c</b>) cross-sectional morphological images.</p>
Full article ">Scheme 1
<p>Schematics of a long-acting injectable alginate hydrogel and the mechanism of gelation time control using glucono-delta-lactone (GDL) and pyridoxal phosphate (PLP).</p>
Full article ">
17 pages, 7803 KiB  
Article
Effect of Adding Gold Nanoparticles on the Anti-Candidal Activity and Release Profile of Itraconazole from Hydrogels
by Radosław Balwierz, Paweł Biernat, Dawid Bursy, Mariia Shanaida, Katarzyna Kasperkiewicz, Agata Jasińska-Balwierz and Wioletta Ochędzan-Siodłak
Appl. Sci. 2024, 14(23), 11125; https://doi.org/10.3390/app142311125 - 29 Nov 2024
Viewed by 552
Abstract
Gold nanoparticles have been identified as a promising avenue for the development of drug carriers, particularly in the context of antimicrobial drug delivery, where limited solubility represents a significant challenge. The ability of gold nanoparticles to penetrate biofilms and disrupt fungal cell membranes [...] Read more.
Gold nanoparticles have been identified as a promising avenue for the development of drug carriers, particularly in the context of antimicrobial drug delivery, where limited solubility represents a significant challenge. The ability of gold nanoparticles to penetrate biofilms and disrupt fungal cell membranes makes them an effective tool to support antifungal therapy, especially against resistant strains. Gold nanoparticles also demonstrate synergistic effects with chemotherapeutics and can influence the release profile of the active substances. This study aimed to develop a topical hydrogel drug formulation containing itraconazole (ITZ), with the addition of gold nanoparticles, to enhance its therapeutic properties. Due to ITZ’s poor water solubility, three types of the gold nanoparticles (AuNPs) of different sizes were synthesized and subsequently coated with itraconazole. The resulting formulations were incorporated into carbopol gels and their ability to diffuse through semipermeable membranes was assessed. The findings demonstrated that the combination of gold nanoparticles and itraconazole elevated the diffusion coefficient to twice the level observed in gels without nanoparticles. Furthermore, the combined effect of gold nanoparticles and itraconazole against a reference Candida albicans strain was investigated. The combination of gold nanoparticles and itraconazole demonstrated a growth-inhibitory effect on this strain, indicating that this formulation could potentially be employed in the treatment of fungal infections. The study confirms that hydrogels with itraconazole and gold nanoparticles can be obtained, offering enhanced drug diffusion. Full article
(This article belongs to the Special Issue Nanomaterials in Medical Diagnosis and Therapy)
Show Figures

Figure 1

Figure 1
<p>The FTIR spectra of the studied materials.</p>
Full article ">Figure 2
<p>Dependence of the diffusion rate on the time element for the obtained hydrogel formulations.</p>
Full article ">Figure 3
<p>Amount of active substance released as a function of time element for all gel formulations.</p>
Full article ">Figure 4
<p><span class="html-italic">Candida albicans</span> growth kinetics tested for selected hydrogels formulations.</p>
Full article ">
16 pages, 2157 KiB  
Article
A Fluorescent Perspective on Water Structuring: ACDAN in Salt Solutions and Hydrogels
by Giuseppe De Luca, Vittorio Ferrara, Bruno Pignataro, Valeria Vetri and Giuseppe Sancataldo
Biophysica 2024, 4(4), 619-633; https://doi.org/10.3390/biophysica4040041 (registering DOI) - 28 Nov 2024
Viewed by 221
Abstract
The interactions and structural organization of water molecules play a crucial role in a wide range of physical, chemical, and biological processes. The ability of water to form hydrogen bonds (H-bonds) underpins its unique properties and enables it to respond dynamically to various [...] Read more.
The interactions and structural organization of water molecules play a crucial role in a wide range of physical, chemical, and biological processes. The ability of water to form hydrogen bonds (H-bonds) underpins its unique properties and enables it to respond dynamically to various environmental factors. These interactions at the molecular level may affect vital processes like protein folding, enzyme activity, and cellular organization. The presence of solutes and spatial constraints can alter the H-bonding network of water, and these effects are ubiquitous in the biological environment. In this study, we analyzed the fluorescence of 2-acetyl-6-(dimethylamino)naphthalene (ACDAN) fluorescence emission in water solutions containing kosmotropic and chaotropic salts and in agar hydrogels. Recently, this dye has proven invaluable in studying water network structure and dynamics, as its fluorescence signal changes based on the local dielectric environment, revealing variations in the dipolar relaxation of water. Our results show that ACDAN spectral response correlates with the degree of water ordering, providing important insights into solute–water interactions and water dynamics in free and confined environments. Full article
(This article belongs to the Special Issue Biomedical Optics 2.0)
Show Figures

Figure 1

Figure 1
<p>NIR absorbance of the water solution of kosmotropic salts of (<b>a</b>) sodium dihydrogen phosphate (Na<sub>2</sub>HPO<sub>4</sub>) and (<b>b</b>) sodium sulfate (Na<sub>2</sub>SO<sub>4</sub>) at 1 M, 2 M, and 3 M. The spectrum of pure water is reported for comparison. NIR spectra have been normalized by the water mass concentration. ACDAN fluorescence emission spectra in (<b>c</b>) Na<sub>2</sub>HPO<sub>4</sub> and (<b>d</b>) Na<sub>2</sub>SO<sub>4</sub> water solutions (λ<sub>exc</sub> = 380 nm, detection range = 370–650 nm). Fluorescence spectra have been normalized to the maximum intensity.</p>
Full article ">Figure 2
<p>NIR absorbance of the water solution of chaotropic salts of (<b>a</b>) sodium chloride (NaCl) and (<b>b</b>) sodium perchlorate (NaClO<sub>4</sub>) at 1 M, 2 M, and 3 M. The spectrum of pure water is reported for comparison. NIR spectra have been normalized by the water mass concentration. ACDAN fluorescence emission spectra in (<b>c</b>) NaCl and (<b>d</b>) NaClO<sub>4</sub> water solutions (λ<sub>exc</sub> = 380 nm, detection range = 370–650 nm). Fluorescence spectra have been normalized to the maximum intensity.</p>
Full article ">Figure 3
<p>(<b>a</b>) ACDAN fluorescence emission spectra (λ<sub>exc</sub> = 380 nm, detection range = 370–650 nm) measured in agar hydrogels as a function of agar concentration ranging from 0.01% to 1.0% (<span class="html-italic">wt</span>/<span class="html-italic">wt</span>). ACDAN fluorescence emission spectrum in water is reported as well. Fluorescence spectra have been normalized to the maximum intensity. (<b>b</b>) GP analysis of the ACDAN fluorescence emission band, measured at 475 nm and 550 nm. The empty green circle is the GP value calculated from the water spectrum.</p>
Full article ">Figure 4
<p>Time evolution of hydrogel formation was monitored on a sample consisting of a 0.5% <span class="html-italic">wt</span>/<span class="html-italic">wt</span> agar solution for 250 min. (<b>a</b>) Rayleigh scattering of the sample measured at 650 nm during 20 °C thermal gelification. (<b>b</b>) ACDAN fluorescence emission spectra as a function of time (λ<sub>exc</sub> = 380 nm, detection range = 370–650 nm). Fluorescence spectra have been normalized to the maximum intensity. (<b>c</b>) GP analysis of the ACDAN fluorescence emission band, calculated at 475 nm and 550 nm as a function of time. (<b>d</b>) Differential spectra of the 1450 nm NIR absorption band. (<b>e</b>) Maximum of the differential NIR spectra as a function of time.</p>
Full article ">
18 pages, 980 KiB  
Article
3D Bioprintable Self-Healing Hyaluronic Acid Hydrogel with Cysteamine Grafting for Tissue Engineering
by Kasula Nagaraja, Amitava Bhattacharyya, Minsik Jung, Dajeong Kim, Mst Rita Khatun and Insup Noh
Gels 2024, 10(12), 780; https://doi.org/10.3390/gels10120780 (registering DOI) - 28 Nov 2024
Viewed by 236
Abstract
The abundance of hyaluronic acid (HA) in human tissues attracts its thorough research in tissue regenerating scaffolds and 3D bioprintable hydrogel preparation. Though methacrylation of HA can lead to photo-crosslinkable hydrogels, the catalyst has toxicity concerns, and the hydrogel is not suitable for [...] Read more.
The abundance of hyaluronic acid (HA) in human tissues attracts its thorough research in tissue regenerating scaffolds and 3D bioprintable hydrogel preparation. Though methacrylation of HA can lead to photo-crosslinkable hydrogels, the catalyst has toxicity concerns, and the hydrogel is not suitable for creating stable complex 3D structures using extrusion 3D bioprinting. In this study, a dual crosslinking on methacrylated HA is introduced, using cysteamine-grafted HA and varying concentrations of 2-hydroxy ethyl acrylate. The resultant hydrogel is suitable for extrusion 3D printing (or bioprinting), mechanically robust, self-standing, stable in phosphate-buffered saline at 37 °C for more than 42 days, has high water absorption capacity with a low swelling ratio (1.5), and exhibits self-healing and adhesive properties. Complex 3D structures like ears and pyramid shapes with more than 2 cm of height are 3D printed using the optimized composition. All the synthesized hydrogels have shown nontoxicity and cell-supportiveness. Loading of cells, tetracycline, and bovine serum albumin into the hydrogel led to better bioink properties such as cell attachment, growth, and proliferation for osteoblast cells. The test results suggest that this hydrogel is biocompatible and has potential for 3D bioprinting of self-standing structures in bioink form in tissue engineering and regenerative medicine. Full article
(This article belongs to the Special Issue Recent Trends in Gels for 3D Printing)
16 pages, 3008 KiB  
Article
Adsorption of Cr(VI) Using Organoclay/Alginate Hydrogel Beads and Their Application to Tannery Effluent
by Mayra X. Muñoz-Martinez, Iván F. Macías-Quiroga and Nancy R. Sanabria-González
Gels 2024, 10(12), 779; https://doi.org/10.3390/gels10120779 - 28 Nov 2024
Viewed by 329
Abstract
The tanning industry is among the most environmentally harmful activities globally due to the pollution of lakes and rivers from its effluents. Hexavalent chromium, a metal in tannery effluents, has adverse effects on human health and ecosystems, requiring the development of removal techniques. [...] Read more.
The tanning industry is among the most environmentally harmful activities globally due to the pollution of lakes and rivers from its effluents. Hexavalent chromium, a metal in tannery effluents, has adverse effects on human health and ecosystems, requiring the development of removal techniques. This study assessed the efficacy of organobentonite/alginate hydrogel beads in removing Cr(VI) from a fixed-bed adsorption column system. The synthesized organobentonite (OBent) was encapsulated in alginate, utilizing calcium chloride as a crosslinking agent to generate hydrogel beads. The effects of the volumetric flow rate, bed height, and initial Cr(VI) concentration on a synthetic sample were analyzed in the experiments in fixed-bed columns. The fractal-like modified Thomas model showed a good fit to the experimental data for the asymmetric breakthrough curves, confirmed by the high R2 correlation coefficients and low χ2 values. The application of organoclay/alginate hydrogel beads was confirmed with a wastewater sample from an artisanal tannery industry in Belén (Nariño, Colombia), in which a Cr(VI) removal greater than 99.81% was achieved. Organobentonite/alginate hydrogels offer the additional advantage of being composed of a biodegradable polymer (sodium alginate) and a natural material (bentonite-type clay), resulting in promising adsorbents for the removal of Cr(VI) from aqueous solutions in both synthetic and real water samples. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Color changes of the prepared hydrogel beads with different organoclay concentrations.</p>
Full article ">Figure 2
<p>X-ray diffraction patterns of alginate, sodium bentonite, organobentonite, and the hydrogel beads with 2 wt.% organoclay.</p>
Full article ">Figure 3
<p>Size distribution (diameter) of the organobentonite/alginate hydrogel beads.</p>
Full article ">Figure 4
<p>Effect of the organobentonite concentration in the hydrogel beads on Cr(VI) removal. Conditions: [Cr(VI)] = 25 mg/L, V = 50 mL, pH = 3.4, stirring speed = 150 rpm, organobentonite mass in hydrogel beads = 220 ± 3 mg, T = 20 ± 1 °C.</p>
Full article ">Figure 5
<p>Effect of pH on Cr(VI) removal by OBent (2%)/Alg hydrogel beads. Conditions: [Cr(VI)] = 25 mg/L, V = 50 mL, stirring speed = 150 rpm, organobentonite mass in the hydrogel beads = 220 ± 3 mg, T = 20 ± 1 °C.</p>
Full article ">Figure 6
<p>Breakthrough curves for Cr(VI) adsorption on OBent (2%)/Alg at different conditions (Dash line + symbol) and fits to the fractal-like modified Thomas model. (<b>a</b>) Effect of the volumetric flow rate, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> = 20 mg/L, pH = 3.4, and <math display="inline"><semantics> <mrow> <mi>h</mi> </mrow> </semantics></math> = 10 cm; (<b>b</b>) effect of the bed height, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mn>0</mn> </mrow> </msub> </mrow> </semantics></math> = 20 mg/L, pH = 3.4, and <math display="inline"><semantics> <mrow> <mi>Q</mi> </mrow> </semantics></math> = 3 mL/min; and (<b>c</b>) effect of the initial Cr(VI) concentration, pH = 3.4, <math display="inline"><semantics> <mrow> <mi>Q</mi> </mrow> </semantics></math> = 3.0 mL/min, and h = 15 cm.</p>
Full article ">Figure 7
<p>(<b>a</b>) Leather tanning tank; (<b>b</b>) sample of tannery wastewater.</p>
Full article ">Figure 8
<p>Successive Cr(VI) adsorption cycles on OBent (2%)/Alg hydrogel beads.</p>
Full article ">Figure 9
<p>Appearance of hydrogel beads. (<b>a</b>) Before and after Cr(III) removal; (<b>b</b>) before and after Cr(VI) removal.</p>
Full article ">
29 pages, 6429 KiB  
Article
Investigation of Silver- and Plant Extract-Infused Polymer Systems: Antioxidant Properties and Kinetic Release
by Magdalena Bańkosz and Bożena Tyliszczak
Int. J. Mol. Sci. 2024, 25(23), 12816; https://doi.org/10.3390/ijms252312816 - 28 Nov 2024
Viewed by 218
Abstract
This study evaluated the impact of silver particles, suspended in Arnica montana flower extract, on the physicochemical characteristics and release dynamics of antioxidant compounds in PVP (polyvinylpyrrolidone)-based hydrogel systems. The hydrogels were synthesized via photopolymerization with fixed amounts of crosslinker (PEGDA) and photoinitiator, [...] Read more.
This study evaluated the impact of silver particles, suspended in Arnica montana flower extract, on the physicochemical characteristics and release dynamics of antioxidant compounds in PVP (polyvinylpyrrolidone)-based hydrogel systems. The hydrogels were synthesized via photopolymerization with fixed amounts of crosslinker (PEGDA) and photoinitiator, while the concentration of the silver-infused extract was systematically varied. Key properties, including the density, porosity, surface roughness, swelling capacity, and water vapor transmission rate (WVTR), were quantitatively analyzed. The results demonstrated that increasing the silver content reduced the hydrogel density from 0.6669 g/cm3 to 0.2963 g/cm3 and increased the porosity from 4% to 11.04%. The surface roughness parameters (Ra) rose from 8.42 µm to 16.33 µm, while the WVTR increased significantly from 65.169 g/m2·h to 93.772 g/m2·h. These structural changes directly influenced the release kinetics of antioxidant compounds, with kinetic modeling revealing silver-dependent variations in the evaluated release mechanisms. This innovative approach of integrating silver particles and plant-derived antioxidants into hydrogels highlights a novel pathway for tailoring material properties. The observed enhanced porosity and moisture regulation underscore the hydrogels’ potential for biomedical applications, particularly in wound care, where controlled moisture and antioxidant delivery are critical. These findings provide new insights into how silver particles modulate hydrogel structures and functionalities. Full article
(This article belongs to the Special Issue Structural and Functional Polymer Materials in Biomedicine)
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 371
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 - 28 Nov 2024
Viewed by 229
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 ">
Back to TopTop