[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (306)

Search Parameters:
Keywords = bioadhesives

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 3619 KiB  
Article
Bioadhesive Chitosan Films Loading Curcumin for Safe and Effective Skin Cancer Topical Treatment
by Seila Tolentino, Mylene M. Monteiro, Felipe Saldanha-Araújo, Marcilio Cunha-Filho, Tais Gratieri, Eliete N. Silva Guerra and Guilherme M. Gelfuso
Pharmaceutics 2025, 17(1), 18; https://doi.org/10.3390/pharmaceutics17010018 - 26 Dec 2024
Viewed by 363
Abstract
Background/Objectives: This study aimed to evaluate the safety and efficacy of chitosan-based bioadhesive films for facilitating the topical delivery of curcumin in skin cancer treatment, addressing the pharmacokinetic limitations associated with oral administration. Methods: The films, which incorporated curcumin, were formulated [...] Read more.
Background/Objectives: This study aimed to evaluate the safety and efficacy of chitosan-based bioadhesive films for facilitating the topical delivery of curcumin in skin cancer treatment, addressing the pharmacokinetic limitations associated with oral administration. Methods: The films, which incorporated curcumin, were formulated using varying proportions of chitosan, polyvinyl alcohol, Poloxamer® 407, and propylene glycol. These films were assessed for stability, drug release, in vitro skin permeation, cell viability (with and without radiotherapy), and skin irritation. Results: The films demonstrated physical stability and preserved curcumin content at room temperature for 90 days. Drug release was effectively controlled during the first 8 h, with release rates ranging from 51.6 ± 4.8% to 65.6 ± 13.0%. The films also enhanced drug penetration into the skin compared to a curcumin solution used as a control (stratum corneum: 1.3 ± 0.1 to 1.9 ± 0.8 µg/cm²; deeper skin layers: 1.7 ± 0.1 to 2.7 ± 0.2 µg/cm²). A cytotoxicity test on metastatic melanoma cells showed that curcumin at topical doses exerted activity similar to that delivered via the skin. Furthermore, curcumin alone was more effective in inhibiting tumor cells than radiotherapy alone (p < 0.01), with no additional benefit observed when curcumin was combined with radiotherapy. Finally, irritation tests confirmed that the films were safe for topical application. Conclusion: The developed chitosan-based bioadhesive films represent a promising alternative for the topical treatment of skin tumors using curcumin. Full article
Show Figures

Figure 1

Figure 1
<p>Images captured from the F1, F2, and F3 films at the predefined times for the stability study (0, 7, 30, 60, and 90 days) and variation in curcumin content over time. The storage conditions were room temperature (RT) and 40 °C.</p>
Full article ">Figure 2
<p>Curcumin release profiles from films (F1, F2, and F3) over 24 h, with collections at 1, 2, 4, 6, 8, 18, 20, 22, and 24 h and recording photographs of the films at the end of the study.</p>
Full article ">Figure 3
<p>Curcumin recovered from the skin layers (µg/cm<sup>2</sup>) after a 24 h treatment with the films compared to the control. (<b>a</b>) Stratum corneum, (<b>b</b>) remaining skin. The data represent the mean of 5 determinations ± standard deviation. #, values below the limit of quantification; (*), <span class="html-italic">p</span> ≤ 0.05; and (****), <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 4
<p>Effect of curcumin on the viability of MeWo cell line. The Kruskal–Wallis test was employed for cell viability data, and IC<sub>50</sub> values were calculated following nonlinear regression on dose–response curves. The data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 9). (**), <span class="html-italic">p</span> ≤ 0.01; and (****), <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 5
<p>Effect of curcumin on the viability of MeWo cell line, with or without three different doses of radiotherapy. The data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 9). Statistical analysis was performed using ANOVA. (*), <span class="html-italic">p</span> ≤ 0.05; (**), <span class="html-italic">p</span> ≤ 0.01; and (***); <span class="html-italic">p</span> ≤ 0.001.</p>
Full article ">Figure 6
<p>Illustrative sequence of photographic records taken during the HET-CAM assay demonstrating the effects with positive control (1.0 mol/L NaOH), negative control (PBS), and solutions containing each of the films (F1, F2, and F3) solubilized in water on the chorioallantoic membrane after 30 s, 2 min, and 5 min of application.</p>
Full article ">Figure 7
<p>Effect of chitosan on the viability of the human keratinocytes (HaCaT cell line). The Kruskal–Wallis test was employed for cell viability data. Due to the sustained cell viability, it was not possible to calculate the IC<sub>50</sub> from the obtained dose–response curve. The data are presented as the mean ± standard deviation (<span class="html-italic">n</span> = 9).</p>
Full article ">
15 pages, 7639 KiB  
Article
Superhydrophobic Surfaces as a Potential Skin Coating to Prevent Jellyfish Stings: Inhibition and Anti-Tentacle Adhesion in Nematocysts of Jellyfish Nemopilema nomurai
by Yichen Xie, Yuanyuan Sun, Rongfeng Li, Song Liu, Ronge Xing, Pengcheng Li and Huahua Yu
Materials 2024, 17(23), 5983; https://doi.org/10.3390/ma17235983 - 6 Dec 2024
Viewed by 519
Abstract
The development of skin-protective materials that prevent the adhesion of cnidarian nematocysts and enhance the mechanical strength of these materials is crucial for addressing the issue of jellyfish stings. This study aimed to construct superhydrophobic nanomaterials capable of creating a surface that inhibits [...] Read more.
The development of skin-protective materials that prevent the adhesion of cnidarian nematocysts and enhance the mechanical strength of these materials is crucial for addressing the issue of jellyfish stings. This study aimed to construct superhydrophobic nanomaterials capable of creating a surface that inhibits nematocyst adhesion, therefore preventing jellyfish stings. We investigated wettability and nematocyst adhesion on four different surfaces: gelatin, polydimethylsiloxane (PDMS), dodecyl trichlorosilane (DTS)-modified SiO2, and perfluorooctane triethoxysilane (PFOTS)-modified TiO2. Our findings revealed that an increase in hydrophobicity significantly inhibited nematocyst adhesion. Furthermore, DTS-modified sprayed SiO2 and PFOTS-modified sprated TiO2 were further enhanced with low-surface-energy substances—cellulose nanofibers (CNF) and chitin nanocrystals (ChNCs)—to improve both hydrophobicity and mechanical strength. After incorporating CNF and ChNCs, the surface of s-TiO2-ChNCs exhibited a contact angle of 153.49° even after undergoing abrasion and impact tests, and it maintained its hydrophobic properties with a contact angle of 115.21°. These results indicate that s-TiO2-ChNCs can serve as an effective skin coating to resist tentacle friction. In conclusion, this study underscores the importance of utilizing hydrophobic skin materials to inhibit the adhesion of tentacle nematocysts, providing a novel perspective for protection against jellyfish stings. Full article
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">Nemopilema Nomurai</span> tentacle sample and nematocysts on surface, bar = 100 μm.</p>
Full article ">Figure 2
<p>Abrasion and impact test operation mode. (<b>a</b>) Jellyfish suspension impact test, (<b>b</b>) Jellyfish tentacle impact test.</p>
Full article ">Figure 3
<p>Images of the adhesion behavior of surfaces and morphology of the surface after jellyfish adhesion. (<b>a</b>) Gelatin surface, (<b>b</b>) PDMS surface, (<b>c</b>) PFOTS-modified sprayed TiO<sub>2</sub>, (<b>d</b>) DTS-modified sprayed SiO<sub>2</sub>, (<b>e</b>) Gelatin surface, the red arrows represent the cracked undischarged nematocysts, (<b>f</b>) PDMS surface, (<b>g</b>) PFOTS-modified sprayed TiO<sub>2</sub>, (<b>h</b>) DTS-modified sprayed SiO<sub>2</sub>, bar = 50 μm.</p>
Full article ">Figure 4
<p>The variations in the CA of different CNFs after being impacted by <span class="html-italic">N. nomurai</span> suspension for different durations. (<b>a</b>) Post-impact CA of CNF-low-surface-energy material, (<b>b</b>) CNF-fluorosilane, (<b>c</b>) s-SiO<sub>2</sub>-CNF, (<b>d</b>) s-TiO<sub>2</sub>-CNF.</p>
Full article ">Figure 5
<p>The variations in the CA of different ChNCs after being impacted by <span class="html-italic">N. nomurai</span> suspension for different durations. (<b>a</b>) ChNCs-low-surface-energy material compound post-impact CA, (<b>b</b>) ChNCs-fluorosilane, (<b>c</b>) s-SiO<sub>2</sub>-ChNCs, (<b>d</b>) s-TiO<sub>2</sub>-ChNCs.</p>
Full article ">Figure 6
<p>Seawater immersion test for s-TiO<sub>2</sub>-ChNCs on Bama pig skin, contact angle (CA), sliding angle (SA), contact-angle hysteresis, and images of s-TiO<sub>2</sub>-ChNCs material on the skin surface after seawater immersion from 0 min to 120 min. The NA in the figure indicates that the SA cannot be measured.</p>
Full article ">Figure 7
<p>SEM images of jellyfish tentacle contact on Bama pig skin and s-TiO<sub>2</sub>-ChNCs covered skin. (<b>a</b>) Bama pig skin before tentacle contact, (<b>b</b>) Bama pig skin after tentacle contact, red arrows represent the undischarged nematocysts and tubules, (<b>c</b>) s-TiO<sub>2</sub>-ChNCs covered skin before tentacle contact, (<b>d</b>) s-TiO<sub>2</sub>-ChNCs covered skin after tentacle contact, bar = 100 μm.</p>
Full article ">Figure 8
<p>XPS wide-scan spectra for (a) skin, (b) skin-tentacles, (c) s-TiO<sub>2</sub>-ChNCs, and (d) s-TiO<sub>2</sub>-ChNCs-tentacles.</p>
Full article ">Figure 9
<p>Diagram of the surface contact of the tentacles and s-TiO<sub>2</sub>-ChNCs.</p>
Full article ">
29 pages, 5701 KiB  
Article
Polysaccharide-Stabilized Semisolid Emulsion with Vegetable Oils for Skin Wound Healing: Impact of Composition on Physicochemical and Biological Properties
by Giovanna Araujo de Morais Trindade, Laiene Antunes Alves, Raul Edison Luna Lazo, Kamila Gabrieli Dallabrida, Jéssica Brandão Reolon, Juliana Sartori Bonini, Karine Campos Nunes, Francielle Pelegrin Garcia, Celso Vataru Nakamura, Fabiane Gomes de Moraes Rego, Roberto Pontarolo, Marcel Henrique Marcondes Sari and Luana Mota Ferreira
Pharmaceutics 2024, 16(11), 1426; https://doi.org/10.3390/pharmaceutics16111426 - 8 Nov 2024
Viewed by 779
Abstract
Background/Objectives: The demand for natural-based formulations in chronic wound care has increased, driven by the need for biocompatible, safe, and effective treatments. Natural polysaccharide-based emulsions enriched with vegetable oils present promising benefits for skin repair, offering structural support and protective barriers suitable for [...] Read more.
Background/Objectives: The demand for natural-based formulations in chronic wound care has increased, driven by the need for biocompatible, safe, and effective treatments. Natural polysaccharide-based emulsions enriched with vegetable oils present promising benefits for skin repair, offering structural support and protective barriers suitable for sensitive wound environments. This study aimed to develop and evaluate semisolid polysaccharide-based emulsions for wound healing, incorporating avocado (Persea gratissima) and blackcurrant (Ribes nigrum) oils (AO and BO, respectively). Both gellan gum (GG) and kappa-carrageenan (KC) were used as stabilizers due to their biocompatibility and gel-forming abilities. Methods: Four formulations were prepared (F1-GG-AO; F2-KC-AO; F3-GG-BO; F4-KC-BO) and evaluated for physicochemical properties, spreadability, rheology, antioxidant activity, occlusive and bioadhesion potential, biocompatibility, and wound healing efficacy using an in vitro scratch assay. Results: The pH values (4.74–5.06) were suitable for skin application, and FTIR confirmed excipient compatibility. The formulations showed reduced occlusive potential, pseudoplastic behavior with thixotropy, and adequate spreadability (7.13–8.47 mm2/g). Lower bioadhesion indicated ease of application and removal, enhancing user comfort. Formulations stabilized with KC exhibited superior antioxidant activity (DPPH scavenging) and fibroblast biocompatibility (CC50% 390–589 µg/mL) and were non-hemolytic. Both F2-KC-AO and F4-KC-BO significantly improved in vitro wound healing by promoting cell migration compared to other formulations. Conclusions: These findings underscore the potential of these emulsions for effective wound treatment, providing a foundation for developing skin care products that harness the therapeutic properties of polysaccharides and plant oils in a natural approach to wound care. Full article
(This article belongs to the Special Issue Dosage Form Design and Delivery Therapy for Skin Disorders)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the formulation and characterization procedures. The preparation of the semisolid emulsion involves several sequential steps (<b>A</b>): weighing the individual components for the oil phase (OP) and aqueous phase (AP), heating each phase separately to 70 °C to ensure proper dissolution and mixing, combining the phases by gradually pouring the aqueous phase (AP) into the oil phase (OP) under constant stirring to form a uniform emulsion, and obtaining the final gel–cream formulation. The emulsion was subsequently characterized through various analyses (<b>B</b>): Fourier-transform infrared spectroscopy (FTIR) to assess molecular interactions and confirm compatibility among components, centrifugation to evaluate physical stability and detect any phase separation, spreadability and reology testing to determine ease of application and coverage on the skin, density measurement to assess formulation consistency, pH measurement with a pH meter to ensure suitability for skin application, bioadhesion and occlusion potential, antioxidant activity, cytotoxicity testing using cell cultures to evaluate biocompatibility and potential safety for skin use, and wound healing assay to determine efficacy.</p>
Full article ">Figure 2
<p>Macroscopic (<b>A</b>) and microscopic (<b>B</b>) images of polysaccharide-based semisolid emulsions containing vegetable oils. Overall, the formulations have a whitish color, homogeneous aspect, and shiny texture. The microscopic evaluation indicates that the system effectively dispersed the oil, keeping it stable within the semisolid structure. Abbreviations: GG—Gellan gum; KC—<span class="html-italic">Kappa</span>-carrageenan; BO—Blackcurrant Oil; AO—Avocado Oil.</p>
Full article ">Figure 3
<p>Infrared spectra of raw materials (<b>A</b>) and semisolid emulsions (<b>B</b>). The spectra exhibit characteristic peaks corresponding to the functional groups present in the substances. Additionally, these spectra support the compatibility among the components, as the absence of significant new peaks suggests no chemical interaction altering the molecular structure of the excipients.</p>
Full article ">Figure 4
<p>PCA model. In (<b>A</b>,<b>B</b>) are the eigenvalues graphs, which indicate that these three principal components encompass most of the chemical information in the raw materials. The red circles represent the principal components selected for the model. In (<b>C</b>,<b>D</b>) are the score plot graphs that reveal a distinct differentiation is observable between the formulations containing GG and KC, emphasizing these polysaccharides’ influence on the formulations’ ultimate chemical composition.</p>
Full article ">Figure 5
<p>Spreadability profile (<b>A</b>), spreadability factor (<b>B</b>), and viscosity (<b>C</b>) of semisolid emulsions. The developed emulsions demonstrated an increased spreading area with the application of more weight, suggesting they can expand more easily under pressure. Moreover, rheological measurements supported this behavior, as the complex viscosity (η*) of all formulations decreased with increasing angular frequency, which is a characteristic of pseudoplastic materials.</p>
Full article ">Figure 6
<p>Storage modulus (G′) and loss modulus (G″) as functions of angular frequency (ω). In (<b>A</b>,<b>B</b>) formulations containing AO stabilized with GG and KC, respectively. In (<b>C</b>,<b>D</b>) formulations prepared with BO stabilized with GG and KC, respectively. Data indicates that elastic and viscous behaviors become more pronounced at higher frequencies, suggesting a predominantly elastic rather than viscous behavior.</p>
Full article ">Figure 7
<p>Thixotropy evaluation of F1-GG-AO (<b>A</b>), F2-KC-AO (<b>B</b>), F3-GG-BO (<b>C</b>), and (<b>D</b>) F4-KC-BO. The data show that the material’s structure is temporarily disrupted under shear, but it recovers gradually when the shear is removed, which is characteristic of thixotropic materials.</p>
Full article ">Figure 8
<p>Antioxidant activity. The @ denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) between formulations and their respective blank forms (F1-GG-AO versusF5-GG-B, and F3-GG-BO versus F5-GG-B); # represents the significant difference (<span class="html-italic">p</span> &lt; 0.05) between polysaccharides (F1-GG-AO versus F2-KC-AO, and F5-GG-B versus F6-KC-B). NS means “not significant”. Both oils significantly enhanced the antioxidant potential of GG emulsions compared to the placebo semisolid, while emulsions stabilized with KC demonstrated higher antioxidant properties than those stabilized with GG.</p>
Full article ">Figure 9
<p>Occlusion potential. The @ denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) between formulations and their respective blank forms (F2-KC-AO versusF6-KC-B, and F4-KC-BO versus F6-KC-B); # represents the significant difference (<span class="html-italic">p</span> &lt; 0.05) between polysaccharides (F5-GG-B versus F6-KC-B). NS means “not significant”. Similar occlusion potential was observed among the formulations. Data also suggests that the oily components may negatively affect the KC formulations.</p>
Full article ">Figure 10
<p>Bioadhesion potential in intact and injured skin. The @ denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) between formulations and their respective blank forms; # represents the significant difference (<span class="html-italic">p</span> &lt; 0.05) between polysaccharides with the same oil (F1-GG-AO versusF2-KC-AO, or F3-GG-BO versus F4-KC-BO); * denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) between oils with the same polysaccharide (F1-GG-AO versus F3-GG-BO or F2-KC-AO versus F4-KC-BO); and <span>$</span> represents the significant difference (<span class="html-italic">p</span> &lt; 0.05) between intact and injured skin. NS means “not significant”. All formulations presented significantly higher bioadhesion in intact skin than in injured skin.</p>
Full article ">Figure 11
<p>Effect of F1-GG-AO (<b>A</b>), F2-KC-AO (<b>B</b>), F3-GG-BO (<b>C</b>), F4-KC-BO (<b>D</b>), F5-GG-B (<b>E</b>), and F6-KC-B (<b>F</b>) (1–1000 µg/mL) on the viability of L-929 cells by MTT assay. A negative control (non–treated cells) was conducted and considered 100% viability. Mean values were calculated from 3 independent results. The * denotes the significative difference from the negative control (<span class="html-italic">p</span> &lt; 0.05). NS means “not significant”. In all formulations examined, the viability of cells is observed to decline as the concentration increases.</p>
Full article ">Figure 12
<p>Hemolytic assay of KC semisolid emulsions. The results showed a hemolytic potential of less than 1% for all tested concentrations of the KC-based emulsions.</p>
Full article ">Figure 13
<p>Representative images showing the progression of healing over time (<b>A</b>) and percentage of open wound area at different times (0, 6, and 24 h) (<b>B</b>) for the F2-KC-AO, F4-KC-BO, and F6-KC-B, compared to the negative control. The * denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) with time zero in the same group, and # denotes the significant difference (<span class="html-italic">p</span> &lt; 0.05) with negative control at the same time. There is a consistent reduction in the area of open wounds over time, with formulations containing oils exhibiting a more pronounced degree of cell migration, which suggests an effective healing process.</p>
Full article ">
18 pages, 6137 KiB  
Article
Decellularized Macroalgae as Complex Hydrophilic Structures for Skin Tissue Engineering and Drug Delivery
by Andreea Luca, Florina-Daniela Cojocaru, Maria Stella Pascal, Teodora Vlad, Isabella Nacu, Catalina Anisoara Peptu, Maria Butnaru and Liliana Verestiuc
Gels 2024, 10(11), 704; https://doi.org/10.3390/gels10110704 - 31 Oct 2024
Cited by 1 | Viewed by 767
Abstract
Due to their indisputable biocompatibility and abundant source, biopolymers are widely used to prepare hydrogels for skin tissue engineering. Among them, cellulose is a great option for this challenging application due to its increased water retention capacity, mechanical strength, versatility and unlimited availability. [...] Read more.
Due to their indisputable biocompatibility and abundant source, biopolymers are widely used to prepare hydrogels for skin tissue engineering. Among them, cellulose is a great option for this challenging application due to its increased water retention capacity, mechanical strength, versatility and unlimited availability. Since algae are an unexploited source of cellulose, the novelty of this study is the decellularization of two different species, freshly collected from the Black Sea coast, using two different chemical surfactants (sodium dodecyl sulphate and Triton X-100), and characterisation of the resulted complex biopolymeric 3D matrices. The algae nature and decellularization agent significantly influenced the matrices porosity, while the values obtained for the hydration degree included them in hydrogel class. Moreover, their capacity to retain and then controllably release an anti-inflammatory drug, ibuprofen, led us to recommend the obtained structures as drug delivery systems. The decellularized macroalgae hydrogels are bioadhesive and cytocompatible in direct contact with human keratinocytes and represent a great support for cells. Finally, it was noticed that human keratinocytes (HaCaT cell line) adhered and populated the structures during a monitoring period of 14 days. Full article
(This article belongs to the Special Issue Novel Functional Gels for Biomedical Applications)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of the decellularized macroalgae (FTIR).</p>
Full article ">Figure 2
<p>Morphology of the decellularized macroalgae: (<b>A</b>)—red algae SDS; (<b>B</b>)—red algae Triton; (<b>C</b>)—green algae SDS; (<b>D</b>)—green algae Triton (SEM and stereomicroscope images—last column).</p>
Full article ">Figure 3
<p>Ibuprofen release from hydrophilic decellularized macroalgae.</p>
Full article ">Figure 4
<p>Bioadhesion properties of the decellularized macroalgae: (<b>A</b>) detachment force; (<b>B</b>) work of adhesion.</p>
Full article ">Figure 5
<p>Cell viability after 24, 48, and 72 h of HaCaT cell culturing with the decellularized macroalgae.</p>
Full article ">Figure 6
<p>HaCaT cells morphology after 7 days contact with decellularized macroalgae: (<b>A</b>) control; (<b>B</b>) red algae SDS; (<b>C</b>) red algae Triton; (<b>D</b>) green algae SDS; (<b>E</b>) green algae Triton; (<b>F</b>) cultured cells and algae surface.</p>
Full article ">Figure 7
<p>Cell population of the decellularized green and red macroalgae after 7 and 14 days of culture, calcein–AM-stained and brightfield images. The green circle indicates a cluster of cells onto the red algae after 7 days of culture.</p>
Full article ">Figure 8
<p>Decellularization process (created with BioRender.com, accessed on 1 August 2024). Macroscopic and microscopic aspect of macroalgae before (day 1) and after (day 6) decellularization.</p>
Full article ">
14 pages, 1888 KiB  
Article
Implementing the Design of Experiments (DoE) Concept into the Development of Mucoadhesive Tablets Containing Orange Peel Extract as a Potential Concept for the Treatment of Oral Infections
by Magdalena Paczkowska-Walendowska, Tomasz M. Karpiński, Ewa Garbiec, Michał Walendowski and Judyta Cielecka-Piontek
Materials 2024, 17(21), 5234; https://doi.org/10.3390/ma17215234 - 28 Oct 2024
Viewed by 819
Abstract
This study explores for the first time the impact of chitosan (CS) with varying molecular weights (MW), orange peel extract concentration, and hydroxypropyl methylcellulose (HPMC) content on the formulation of buccal tablets for treating oral infections. Utilizing a statistical design of experiments (DoE), [...] Read more.
This study explores for the first time the impact of chitosan (CS) with varying molecular weights (MW), orange peel extract concentration, and hydroxypropyl methylcellulose (HPMC) content on the formulation of buccal tablets for treating oral infections. Utilizing a statistical design of experiments (DoE), nine different formulations were evaluated for mechanical properties, dissolution behavior, mucoadhesion, and biological activity. A formulation with high CS MW, 60% orange peel extract, and 8% HPMC, emerged as the optimal formulation, demonstrating superior tabletability, compressibility, and compactibility. Dissolution studies indicated that hesperidin release followed the Higuchi model, with higher extract content enhancing this phenomenon. Mucoadhesion improved with increased HPMC and CS concentrations, although higher extract content reduced bioadhesion. Biological assays showed that higher extract levels boosted antioxidant activity, while CS primarily contributed to anti-inflammatory effects. The optimized formulation exhibited broad antimicrobial activity against key oral pathogens, surpassing the effectiveness of the individual components. Principal component analysis (PCA) further confirmed the significant influence of extract content on tablet properties. These findings suggest that the optimized tablet formulation holds promise for effective buccal delivery in the treatment of oral infections, warranting further investigation in clinical settings. Full article
Show Figures

Figure 1

Figure 1
<p>ATR-IR spectra for orange peel extract [<a href="#B15-materials-17-05234" class="html-bibr">15</a>], CS [<a href="#B26-materials-17-05234" class="html-bibr">26</a>], and HPMC [<a href="#B8-materials-17-05234" class="html-bibr">8</a>] (<b>a</b>), and formulation F1–F9 (<b>b</b>).</p>
Full article ">Figure 2
<p>Tabletability (<b>a</b>), compressibility (<b>b</b>), and compactibility (<b>c</b>) profiles of tablets produced from systems 1–9.</p>
Full article ">Figure 3
<p>Hesperidin dissolution profiles for tablets F1–F9 for both compression pressures (′ for lower compression pressure and ″ for higher compression pressure).</p>
Full article ">Figure 4
<p>Swelling index for tablets F1–F9 for both compression pressures (′ for lower compression pressure and ″ for higher compression pressure).</p>
Full article ">Figure 5
<p>Component of bioadhesion of formulations F1–F9.</p>
Full article ">Figure 6
<p>Prediction based on effects with positive signs (true density, porosity, dissolution after 6 h expressed in mg, swelling index, component of mucoadhesion) (<b>a</b>) and with negative signs (antioxidant and anti-inflammatory activities) (<b>b</b>).</p>
Full article ">Figure 7
<p>Principal component analysis (PCA) showing the factor loading plot considering true density, porosity, dissolution after 6 h expressed in mg, swelling index, component of mucoadhesion (=Muco), and antioxidant (=DPPH) and anti-inflammatory activities (=Hyal).</p>
Full article ">
29 pages, 9207 KiB  
Article
Arginine-Biofunctionalized Ternary Hydrogel Scaffolds of Carboxymethyl Cellulose–Chitosan–Polyvinyl Alcohol to Deliver Cell Therapy for Wound Healing
by Alexandra A. P. Mansur, Sandhra M. Carvalho, Ramayana M. de M. Brito, Nádia S. V. Capanema, Isabela de B. Duval, Marcelo E. Cardozo, José B. R. Rihs, Gabriela G. M. Lemos, Letícia C. D. Lima, Marina P. dos Reys, Ana P. H. Rodrigues, Luiz C. A. Oliveira, Marcos Augusto de Sá, Geovanni D. Cassali, Lilian L. Bueno, Ricardo T. Fujiwara, Zelia I. P. Lobato and Herman S. Mansur
Gels 2024, 10(11), 679; https://doi.org/10.3390/gels10110679 - 23 Oct 2024
Viewed by 1126
Abstract
Wound healing is important for skin after deep injuries or burns, which can lead to hospitalization, long-term morbidity, and mortality. In this field, tissue-engineered skin substitutes have therapy potential to assist in the treatment of acute and chronic skin wounds, where many requirements [...] Read more.
Wound healing is important for skin after deep injuries or burns, which can lead to hospitalization, long-term morbidity, and mortality. In this field, tissue-engineered skin substitutes have therapy potential to assist in the treatment of acute and chronic skin wounds, where many requirements are still unmet. Hence, in this study, a novel type of biocompatible ternary polymer hybrid hydrogel scaffold was designed and produced through an entirely eco-friendly aqueous process composed of carboxymethyl cellulose, chitosan, and polyvinyl alcohol and chemically cross-linked by citric acid, forming three-dimensional (3D) matrices, which were biofunctionalized with L-arginine (L-Arg) to enhance cellular adhesion. They were applied as bilayer skin biomimetic substitutes based on human-derived cell cultures of fibroblasts and keratinocytes were seeded and grown into their 3D porous structures, producing cell-based bio-responsive hybrid hydrogel scaffolds to assist the wound healing process. The results demonstrated that hydrophilic hybrid cross-linked networks were formed via esterification reactions with the 3D porous microarchitecture promoted by foam templating and freeze-drying. These hybrids presented chemical stability, physicochemical properties, high moisture adsorption capacity, surface properties, and a highly interconnected 3D porous structure well suited for use as a skin substitute in wound healing. Additionally, the surface biofunctionalization of these 3D hydrogel scaffolds with L-arginine through amide bonds had significantly enhanced cellular attachment and proliferation of fibroblast and keratinocyte cultures. Hence, the in vivo results using Hairless mouse models (an immunocompromised strain) confirmed that these responsive bio-hybrid hydrogel scaffolds possess hemocompatibility, bioadhesion, biocompatibility, adhesiveness, biodegradability, and non-inflammatory behavior and are capable of assisting the skin wound healing process. 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>) An illustrative representation of a cross-section of the skin, indicating the layers (epidermis and dermis) that were resembled by the 3D hybrid hydrogel, depicted with information regarding its components, chemical formula, and structure. (<b>B</b>) Scheme of the co-culture strategy of the cell therapy product simulating the skin layers, with the deposition of keratinocytes in the membrane layer (resembling basement membrane of skin) and populating the porous layer with fibroblasts to obtain a cellular biomimetic dermal–epidermal skin substitute construct (not to scale).</p>
Full article ">Figure 2
<p>NMR spectra of (<b>A</b>) CMC (<sup>1</sup>H NMR), (<b>B</b>) CHI (<sup>1</sup>H NMR), and (<b>C</b>) PVA ((a) <sup>1</sup>H NMR and (b) <sup>13</sup>C NMR). (<b>D</b>) FTIR spectra of the three polymer components of hydrogels (a) CMC, (b) CHI, and (c) PVA.</p>
Full article ">Figure 3
<p>Results of (<b>A</b>) swelling degree (SD) and (<b>B</b>) gel fraction (GF) for 3D structures produced during different processing stages. Sample identification: REF refers to sample without citric acid (CMC:PVA:CHI); CA is the sample chemically cross-linked with citric acid (CMC:PVA:CHI_CA); FD is the cross-linked ternary hybrid hydrogel after the freeze-drying process (CMC:PVA:CHI_FD); and L-Arg is the cross-linked and freeze-dried sample after biofunctionalization with L-arginine (CMC:PVA:CHI_L-Arg).</p>
Full article ">Figure 4
<p>(<b>A</b>) FTIR spectra (range 2000–750 cm<sup>−1</sup>) for the samples: (a) CMC:PVA:CHI (REF), (b) CMC:PVA:CHI_CA, (c) CMC:PVA:CHI_FD, and (d) CMC:PVA:CHI_L-Arg. (<b>B</b>) XPS spectra of N 1s region for hydrogels: (a) CMC:PVA:CHI_FD and (b) CMC:PVA:CHI_L-Arg. (<b>C</b>) Surface wettability assessment of the 3D hydrogel CMC:PVA:CHI_L-Arg through contact angle measurement. (<b>D</b>) PZC studies for the samples: (a) CMC:PVA:CHI_FD and (b) CMC:PVA:CHI_L-Arg.</p>
Full article ">Figure 5
<p>Digital images (red background), stereoscopic microscope images (SM, 20×), and scanning electron microscopy images (SEM, 50×–100×) of the (<b>A</b>) upper and (<b>B</b>) lower surfaces of the porous three-dimensional structure. (<b>C</b>) SEM image (200×) of cross-section (Section A-A in the detail of the digital image). (<b>D</b>) Volumetric pore size distribution by micro-CT.</p>
Full article ">Figure 6
<p>(<b>A</b>) Results of cell viability (MTT cytotoxicity test) and (<b>B</b>) cell proliferation (measured using resazurin protocol) biological assays for the porous three-dimensional structures (3d = 3 days). (<b>C</b>) Calcein AM (viable cells, green fluorescence, left side images, identified as LIVE) and ethidium homodimer (dead cells, red fluorescence, right side images, identified as DEAD) staining for keratinocyte and fibroblast monoculture cells deposited onto the lower and upper faces of the L-arginine functionalized 3D constructs, respectively (scale bar = 100 µm). Details: schematic representation of cell deposition (not to scale).</p>
Full article ">Figure 7
<p>Fluorescence microscopy images showing representative images of the 3D CMC:PVA:CHI_L-Arg structure, where a co-culture of (<b>C</b>) keratinocytes (upper face—membrane) and (<b>D</b>,<b>E</b>) fibroblasts (lower face—porous structure) were seeded compared to controls ((<b>A</b>) Keratinocyte +Control and (<b>B</b>) Fibroblast +Control). Cells were labeled with calcein AM (viable cells, green fluorescence, left column, identified as LIVE) and ethidium homodimer (dead cells, red fluorescence, right column, identified as DEAD). Scale bar = 100 µm. Detail: schematic representation of cell therapy dermal–epidermal skin substitute (not to scale).</p>
Full article ">Figure 8
<p>(<b>A</b>) Monitoring of the body weight of the mice after inducing skin lesions and applying a cell membrane (CM group) compared to the control group (without membrane). (<b>B</b>) Measurement of the wound area over 14 days for both the cell membrane and control groups. (<b>C</b>) Representative photos showing the wound healing process during the days of the experiment and (<b>D</b>) percentage of wound healing (WHR, %). The control group consisted of healthy female <span class="html-italic">Hairless</span> mice (n = 8), with their wound covered by a “Blood stop” bandaid but with no membrane applied. The CM group included <span class="html-italic">Hairless</span> mice with wounds covered by the tested cell membrane (CM, n = 8). In the graph plot D, (*) indicates statistical differences where all groups on days D06, D10, and D14 were compared to the control group on day D02. (#) Indicates statistical difference where all groups on days D06, D10, and D14 were compared to the membrane-treated (CM) group on day D02. The red asterisk represents the statistical difference between the membrane-treated groups (CM) at D06 and D10. Data are shown as mean ± standard deviation. **/## <span class="html-italic">p</span> &lt; 0.01; ****/#### <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 9
<p>Histopathological analysis of skin lesions in <span class="html-italic">Hairless</span> mice, showing representative images of (<b>A</b>) skin from the control group (no hydrogel applied) and (<b>B</b>) skin from the group treated with the 3D bilayer construct containing fibroblasts and keratinocytes (CM group). Black asterisk: intact epidermis; red asterisk: denser collagen; green asterisk: looser collagen; black arrow: reactive fibroblasts; red arrow: blood capillaries. The larger image displays the lesion area at 20× magnification, with a zoomed-in view of the highlighted area at 40× magnification.</p>
Full article ">
18 pages, 13822 KiB  
Article
Influence of Nitride Coatings on Corrosion Resistance and the Biocompatibility of Titanium Alloy Products
by Catherine Sotova, Oleg Yanushevich, Natella Krikheli, Olga Kramar, Alexey Vereschaka, Semen Shehtman, Filipp Milovich, Valery Zhylinski, Anton Seleznev and Pavel Peretyagin
Metals 2024, 14(11), 1200; https://doi.org/10.3390/met14111200 - 22 Oct 2024
Viewed by 878
Abstract
The bioadhesion of bacteria to the surface of samples with Ti–TiN, Zr–ZrN, Zr–(Zr, Nb)N, and Zr–(Zr, Hf)N coatings was studied via incubation with gram-positive strains of Staphylococcus aureus. The samples were kept at 25 °C for 30 days in a 3% NaCl [...] Read more.
The bioadhesion of bacteria to the surface of samples with Ti–TiN, Zr–ZrN, Zr–(Zr, Nb)N, and Zr–(Zr, Hf)N coatings was studied via incubation with gram-positive strains of Staphylococcus aureus. The samples were kept at 25 °C for 30 days in a 3% NaCl solution. The deposition of coatings slows, whereas oxidation processes intensify. The oxygen content on the TiN and (Zr, Nb)N coating surfaces was higher than that of the Ti sample without a coating. Samples with ZrN and, especially, (Zr, Hf)N coatings resist oxidation better. Regarding bioactivity toward S. aureus, the highest density of biological forms was observed on the surfaces of TiN and (Zr, Hf)N coatings. The lowest density was on the surfaces of uncoated, ZrN-coated, and (Zr, Nb)N-coated samples. On Ti–TiN, Zr–ZrN, and Zr–(Zr, Nb)N coatings, the formation of surface biostructures of a filamentary type was observed. In the uncoated sample, the biostructures have an island character, and in the sample with a Zr–(Zr, Hf)N coating, the formation of extensive areas of biostructures was observed. Between the biostructures and coating, a layer 5 to 15 nm thick was observed, presumably associated with bacterial adhesion. The presence of biostructures on the coating surface can activate or slow oxidation processes. Full article
Show Figures

Figure 1

Figure 1
<p>Mapping the distribution of oxygen and carbon (a marker of biological formations) on the surface of the studied coatings.</p>
Full article ">Figure 2
<p>Analysis of the content of elements in the putative biological structures on the surface of the uncoated Ti–6Al–4V substrate, after corrosion tests in a 3% NaCl solution with preliminary planting of <span class="html-italic">Staphylococcus aureus</span>.</p>
Full article ">Figure 3
<p>Comparison of the average oxygen content (at.%) on the surface of the studied samples with various coatings.</p>
Full article ">Figure 4
<p>Structure of the Ti–TiN coating on Ti–6Al–4V alloy after corrosion testing in a 3% NaCl solution at 25 °C for 720 h: (<b>a</b>) view of the sample surface, where the green line indicates the location of the lamella cutout, (<b>b</b>) general view of the lamella across the coating thickness, and (<b>c</b>) biostructure and its adhesion to the coating.</p>
Full article ">Figure 5
<p>Distribution of elements in the surface layers of the Ti–TiN coating on Ti–6Al–4V alloy at the boundary with the biological structure after corrosion tests in a 3% NaCl solution at 25 °C for 720 h: (<b>a</b>) oxygen content in the coating surface layers and (<b>b</b>) analysis of the elemental composition of the biostructure.</p>
Full article ">Figure 6
<p>Structure of the Zr–ZrN coating on Ti–6Al–4V alloy after corrosion testing in a 3% NaCl solution at 25 °C for 720 h: (<b>a</b>) view of the sample surface, where the green line indicates the location of the lamella cutout and (<b>b</b>) the structure of the coating and biostructure on its surface.</p>
Full article ">Figure 7
<p>Oxygen (O) content in the surface layers of the Zr–ZrN coating on Ti–6Al–4V alloy after corrosion testing in 3% NaCl at 25 °C for 720 h: (<b>a</b>) O content in the coating surface layer, (<b>b</b>) study of the elemental composition of the biostructure, and (<b>c</b>) O content across the coating thickness in the area under the biostructure.</p>
Full article ">Figure 8
<p>(<b>a</b>) Surface of the Zr–(Zr, Nb)N coating with filiform biostructures and the location of the lamella cut out (green line). (<b>b</b>) Section and structure of the biostructure on the coating surface. (<b>c</b>) Structure of the intermediate layer between the biostructure and coating. (<b>d</b>) Distribution of oxygen by the depth of the Zr–(Zr, Nb)N coating on Ti–6Al–4V alloy after corrosion testing in a 3% NaCl solution at 25 °C for 720 h.</p>
Full article ">Figure 9
<p>(<b>a</b>) Surface of the Zr–(Zr, Hf)N coating with biostructures and the location of the lamella cut out (green line). (<b>b</b>) Section and structure of the biostructure on the coating surface. (<b>c</b>) Elemental and phase composition of the biostructure. (<b>d</b>) Distribution of oxygen by the depth of the Zr–(Zr, Hf)N coating on Ti–6Al–4V alloy after corrosion testing in a 3% NaCl solution at 25 °C for 720 h in region <span class="html-italic">A</span> under the biostructure and (<b>e</b>) under the free surface of the coating.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) Surface of the Zr–(Zr, Hf)N coating with biostructures and the location of the lamella cut out (green line). (<b>b</b>) Section and structure of the biostructure on the coating surface. (<b>c</b>) Elemental and phase composition of the biostructure. (<b>d</b>) Distribution of oxygen by the depth of the Zr–(Zr, Hf)N coating on Ti–6Al–4V alloy after corrosion testing in a 3% NaCl solution at 25 °C for 720 h in region <span class="html-italic">A</span> under the biostructure and (<b>e</b>) under the free surface of the coating.</p>
Full article ">
19 pages, 3366 KiB  
Article
The Design of Novel 3D-Printed, Moulded, and Oral Viscous Budesonide Formulations for Paediatrics: A Comparative Evaluation of Their Mucoadhesive Properties
by María Magariños-Triviño, Eduardo Díaz-Torres, Javier Suárez-González, Ana Santoveña-Estévez and José B. Fariña
Pharmaceutics 2024, 16(10), 1338; https://doi.org/10.3390/pharmaceutics16101338 - 18 Oct 2024
Viewed by 1048
Abstract
Background/Objectives: Paediatric eosinophilic oesophagitis (EoE) treatment is challenging due to the limited number of age-appropriate formulations. This study aims to develop and evaluate oral viscous suspensions and solid formulations of budesonide (BUD), focusing on their in vitro mucoadhesive properties, to enhance drug delivery [...] Read more.
Background/Objectives: Paediatric eosinophilic oesophagitis (EoE) treatment is challenging due to the limited number of age-appropriate formulations. This study aims to develop and evaluate oral viscous suspensions and solid formulations of budesonide (BUD), focusing on their in vitro mucoadhesive properties, to enhance drug delivery and therapeutic outcomes in paediatric EoE. Methods: This study encompasses the development of oral viscous suspensions and orodispersible solid formulations (moulded tablets and 3D-printed dosage forms) containing BUD. The formulations underwent quality control tests as per the European Pharmacopoeia, chemical stability assessments, and an in vitro evaluation of their mucoadhesiveness properties. Results: A validated analytical method enabled accurate BUD quantification and efficient extraction, and all developed formulations demonstrated chemical stability for 30 days, meeting Ph. Eur. quality standards. Three-dimensional printing using SSE successfully produced 1 mg and 0.5 mg BUD printlets, complying with quality tests for conventional tablets. Formulations containing xanthan gum (L2-XG and P1-0.5-XG) exhibited superior mucoadhesive properties. L2-XG showed significantly higher mucoadhesion than L1-MC. Among the solid formulations, P1-0.5-XG demonstrated the highest mucoadhesive properties. Conclusions: This is the first study to develop solid oral dosage forms of BUD at a very low dose, specifically for paediatric use. The results highlight the potential of 3D printing for developing individualised orodispersible BUD formulations with improved bioadhesion for paediatric EoE treatment. The L2-XG formulation and the XG-containing printlets are the most promising formulations in terms of increasing contact time with the oesophageal mucosa, which could translate into improved therapeutic efficacy in this patient population. Full article
(This article belongs to the Special Issue Advanced Pediatric Drug Formulation Strategies)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the texture analyser. Illustration based on Amorós-Galicia et al., 2023 [<a href="#B45-pharmaceutics-16-01338" class="html-bibr">45</a>]. On the right-hand side of the figure, the texture analyser is shown. The upper probe holds the simulated fluid along with the formulation to be evaluated, while a Petri dish containing the oesophageal tissue covered with simulated medium is placed in the lower part. The graph represents a typical force–displacement curve obtained during the bioadhesion test.</p>
Full article ">Figure 2
<p>Boxplot with outliers of the pressure (<b>A</b>) and weight (<b>B</b>) of the printlets developed. • represents cases or rows with values greater than the height of the boxes multiplied by 1.5.</p>
Full article ">Figure 3
<p>Overlaid UHPLC chromatograms of BUD subjected to stress conditions in basic medium (NaOH 0.1 M) at different times (0 min; 10 min; 60 min; and 24 h). Epimer B (22R); Epimer A (22S).</p>
Full article ">Figure 4
<p>Plots of the log of viscosity versus the log of shear rates of L1-MC and L2-XG during stability tests under different storage conditions. (<b>A</b>). 5 °C; (<b>B</b>). 25 °C; and (<b>C</b>). 40 °C.</p>
Full article ">Figure 5
<p>Visual comparison of the shape and size of the moulded tablets (M1) and printlets (P1, P0.5, P0.5-XG) with a size 0 capsule.</p>
Full article ">Figure 6
<p>Comparative analysis of the maximum bioadhesion results for all formulations studied. (<b>A</b>). Average maximum bioadhesion forces for the doses of each formulation studied. (<b>B</b>). Average maximum bioadhesion forces normalized to 1 g of formulation. (<b>C</b>). Average work adhesion, W (mJ), for the doses of each formulation studied. (<b>D</b>). Average work adhesion, W (mJ), normalised to 1 g of formulation.</p>
Full article ">
12 pages, 1307 KiB  
Article
Development and Evaluation of the Acaricidal Activity of Xantan Gum-Based Hydrogel and Polymeric Nanoparticles Containing Achyrocline satureioides Extract
by Rafaela Regina Fantatto, Annelize Rodrigues Gomes, João Vitor Carvalho Constantini, Camila Fernanda Rodero, Marlus Chorilli, Ana Carolina de Souza Chagas, Ana Melero and Rosemeire Cristina Linhari Rodrigues Pietro
Gels 2024, 10(10), 658; https://doi.org/10.3390/gels10100658 - 14 Oct 2024
Viewed by 836
Abstract
The Rhipicephalus microplus tick causes enormous economic losses in livestock farming around the world. Despite several promising studies carried out with plant extracts such as Achyrocline satureioides against this ectoparasite, a major obstacle is related to pharmaceutical presentation forms. There is no study [...] Read more.
The Rhipicephalus microplus tick causes enormous economic losses in livestock farming around the world. Despite several promising studies carried out with plant extracts such as Achyrocline satureioides against this ectoparasite, a major obstacle is related to pharmaceutical presentation forms. There is no study showing xantan gum-based hydrogel and polycaprolactone nanoparticles containing A. satureioides extract against R. microplus larvae. The objective of this study was to incorporate A. satureioides extract to develop a nanoformulation (AScn) and a hydrogel (ASlh) and evaluate them against R. microplus larvae with the purpose of increasing the contact time of the extract with the larvae and improve the effectiveness. The ethanolic extracts were incorporated in polycaprolactone nanoparticles and characterized via analysis of the mean hydrodinamic diameter and polidispersity index. The xanthan gum-based hydrogel formulation was prepared with crude extract of A. satureioides 40 mg/mL, 0.25% xanthan gum, and 8% poloxamer, to determine the bioadhesiveness of the formulation in bovine leather and the flow rate of the formulation in the animal. The results in larvae demonstrated that when evaluated in the form of a hydrogel (ASlh), mortality was higher, with 91.48% mortality at a concentration of 20 mg/mL presenting itself as an interesting alternative for controlling this ectoparasite. Full article
(This article belongs to the Special Issue Recent Advances in Biopolymer Gels)
Show Figures

Figure 1

Figure 1
<p>Chromatogram obtained from <span class="html-italic">A. satureioides</span> (ASb) inflorescences. Analysis conditions: methanol (LC-MS grade) (MeOH): MilliQ water at 1% acetic acid with a gradient of 38:100% over 45 min, 100% MeOH over 10 min and 100:32% over 7 min and a flow of 0.5 mL/min.</p>
Full article ">Figure 2
<p>Quercetin, 3-O-methylquercetin, and kaempferol molecules found in the ethanolic extract of <span class="html-italic">A. satureioides</span> (ChemDraw Ultra 12.0).</p>
Full article ">Figure 3
<p>(<b>a</b>) Strength of bioadhesion of samples on bovine leather. (<b>b</b>) Bioadhesion work of the samples on the leather.</p>
Full article ">Figure 4
<p>Flow distance for hydrogel formulation containing extract (ASlh), blank (xanthan gum and poloxamer), and water in contact with the bovine leather.</p>
Full article ">Figure 5
<p>Test to determine the flow of formulations on bovine leather (Rafaela Fantatto).</p>
Full article ">
16 pages, 2939 KiB  
Article
In Vitro Mucoadhesive Features of Gliadin Nanoparticles Containing Thiamine Hydrochloride
by Silvia Voci, Agnese Gagliardi, Elena Giuliano, Maria Cristina Salvatici, Antonio Procopio and Donato Cosco
Pharmaceutics 2024, 16(10), 1296; https://doi.org/10.3390/pharmaceutics16101296 - 4 Oct 2024
Viewed by 945
Abstract
Background: Gliadins have aroused significant interest in the last decade as suitable biomaterials for food and pharmaceutical applications. In particular, the oral route is the preferred method of administration for gliadin-based formulations, due to the affinity of this biomaterial for the gut mucosa. [...] Read more.
Background: Gliadins have aroused significant interest in the last decade as suitable biomaterials for food and pharmaceutical applications. In particular, the oral route is the preferred method of administration for gliadin-based formulations, due to the affinity of this biomaterial for the gut mucosa. However, up to now, this has been demonstrated only by means of in vivo or ex vivo studies. Methods: This is why, in this study, various in vitro techniques were employed in order to evaluate the ability of polymeric nanoparticles, made up of a commercial grade of the protein and an etheric surfactant, to interact with porcine gastric mucin. The nanosystems were also used for the encapsulation of thiamine hydrochloride, used as a model of a micronutrient. Results: The resulting systems were characterized by a mean diameter of ~160–170 nm, a narrow size distribution when 0.2–0.6 mg/mL of thiamine was used, and an encapsulation efficiency between 30 and 45% of the drug initially employed. The incubation of the gliadin nanosystems with various concentrations of porcine gastric mucin evidenced the ability of the carriers to interact with the mucus glycoprotein, showing a decreased Zeta potential after a 4 h incubation (from ~−30 to −40 mV), while demonstrating that the encapsulation of the drug did not affect its bioadhesive features. Conclusions: Altogether, these data support the conceivable application of gliadin nanoparticles as formulations for the oral administration of bioactive compounds. Full article
(This article belongs to the Section Drug Delivery and Controlled Release)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Mean diameter, (<b>B</b>) size distribution and (<b>C</b>) surface charge of GNPs prepared with 0.1% <span class="html-italic">w</span>/<span class="html-italic">v</span> of Brij O2 and 0.2–0.8 mg/mL of B1. * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001 vs. empty Brij O2 GNPs. Values are average of three different experiments ± standard deviation. TEM micrographs of (<b>D</b>) empty Brij O2 GNPs and (<b>E</b>) nanosystems prepared with 0.6 mg/mL of drug. Scale bar = 100 nm.</p>
Full article ">Figure 2
<p>TSI profiles of Brij O2 GNPs (0.01% <span class="html-italic">w</span>/<span class="html-italic">v</span> of gliadin and 0.1% <span class="html-italic">w</span>/<span class="html-italic">v</span> of Brij O2) prepared with increasing concentrations of B1. Analyses were carried out at room and body temperature for 1 h.</p>
Full article ">Figure 3
<p>(<b>A</b>) Entrapment efficiency and loading capacity of Brij O2 GNPs prepared with different concentrations of B1 (upper panel); (<b>B</b>) release profiles of B1 (0.2–0.8 mg/mL) from Brij O2 GNPs under SGF (pH 1.2) and SIF (pH 6.8), expressed as a function of amount of B1 entrapped and incubation time. Values represent mean of five different experiments ± standard deviation. Error bars, if not shown, are within symbols. (<b>C</b>) FT-IR spectra of B1-loaded Brij O2 GNPs prepared with initial drug concentration of 0.6 mg/mL.</p>
Full article ">Figure 4
<p>Mean sizes and Z-potential values of Brij O2 and B1 Brij O2 GNPs (0.5 mg/mL) incubated at 37 °C with 1 mL of mucin solutions (0.1–0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) as function of time. ** <span class="html-italic">p</span> &lt; 0.001 and * <span class="html-italic">p</span> &lt; 0.05 vs. means sizes and Z-potential values measured before incubation with mucin. Results are reported as mean of three different experiments ± standard deviation.</p>
Full article ">Figure 5
<p>(<b>A</b>) Amount of mucin (0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span>) adsorbed onto empty and B1-loaded GNPs (0.5 mg/mL), expressed as function of incubation time. (<b>B</b>) FT-IR profile of empty and B1-loaded Brij O2 GNPs (0.5 mg/mL) incubated with 0.5% <span class="html-italic">w</span>/<span class="html-italic">v</span> of mucin for 4 h. ** <span class="html-italic">p</span> &lt; 0.001 and * <span class="html-italic">p</span> &lt; 0.05: empty GNPs vs. B1-loaded GNPs.</p>
Full article ">Scheme 1
<p>Schematic representation of the preparation procedure of gliadin nanoparticles.</p>
Full article ">
16 pages, 2103 KiB  
Article
A Novel Surface Passivation Method of Pyrite within Rocks in Underwater Environments to Mitigate Acid Mine Drainage at Its Source
by Lijun Fan, Tiancheng Han, Xianxing Huang, Yixuan Yang, Tao Zhu, Weiwei Zhai, Daoyong Zhang and Xiangliang Pan
Minerals 2024, 14(10), 973; https://doi.org/10.3390/min14100973 - 27 Sep 2024
Viewed by 739
Abstract
Mitigating acid mine drainage (AMD) at its source, specifically within rocks containing pyrite in underwater environments, poses a significant environmental challenge worldwide. Existing passivation techniques are primarily designed for open-air conditions, involving direct contact with coating materials at a solid–liquid interface, making them [...] Read more.
Mitigating acid mine drainage (AMD) at its source, specifically within rocks containing pyrite in underwater environments, poses a significant environmental challenge worldwide. Existing passivation techniques are primarily designed for open-air conditions, involving direct contact with coating materials at a solid–liquid interface, making them ineffective beneath a water barrier. In this study, we introduce a novel passivation method inspired by the design of underwater bio-adhesives. Tannic acid (TA) combined with polyethylene glycol (PEG) was employed to form a hydrophobic film directly on the pyrite surface, overcoming water resistance and addressing the limitations of current techniques. Electrochemical experiments and chemical leaching experiments were conducted to evaluate the oxidation resistance of the passivating films. TA–PEG-coated pyrite exhibited a lower oxidation rate and a higher static contact angle of 126.2°, achieving suppression efficiencies of 71.6% for total Fe release and 68.1% for total S release. A comprehensive characterization approach, including scanning electron microscopy (SEM), Fourier transform infrared spectroscopy (FTIR), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS), was employed to investigate the passivation mechanism. The results of this study may provide new insights into the preparation of simpler and greener passivating agents to suppress pyrite oxidation at its source in underwater environments. Full article
(This article belongs to the Section Environmental Mineralogy and Biogeochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Nyquist plots; (<b>b</b>) equivalent electrical circuit model; (<b>c</b>) Tafel polarization curves; (<b>d</b>) CV curves. All measurements were conducted in a 0.2 M Na<sub>2</sub>SO<sub>4</sub> electrolyte with a pH value of 2.0.</p>
Full article ">Figure 2
<p>Effects of different passivation on the antioxidation ability of pyrite: evolution of total Fe (<b>a</b>) and total S (<b>b</b>) concentrations in leachates of uncoated and a series of coated pyrites over 48 h. Staged Fe (<b>c</b>) and total S (<b>d</b>) concentrations in uncoated pyrite and TA-5.0–PEG-coated pyrite (hereafter referred to as stages I, II, and III, with a duration of 7 days per stage); 100 mL of HCl solution (pH = 1.0) was refreshed between stages. Standard deviations (<span class="html-italic">n</span> = 3) were represented by error bars.</p>
Full article ">Figure 3
<p>SEM images of (<b>a</b>) uncoated pyrite; (<b>b</b>) TA-coated pyrite; (<b>c</b>) TA-0.1–PEG-coated pyrite; (<b>d</b>) TA-1.0–PEG-coated pyrite; (<b>e</b>) TA-2.0–PEG-coated pyrite; and (<b>f</b>) TA-5.0–PEG-coated pyrite.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>f</b>) Static water contact angles of uncoated and coated pyrites.</p>
Full article ">Figure 5
<p>FTIR (<b>a</b>) and Raman spectra (<b>b</b>) of uncoated pyrite, TA-coated pyrite, and TA-5.0–PEG-coated pyrite.</p>
Full article ">Figure 6
<p>XPS spectra of C 1s (<b>a</b>), O 1s (<b>b</b>), and Fe 2p (<b>c</b>) of uncoated pyrite, TA-coated pyrite, and TA-5.0–PEG-coated pyrite.</p>
Full article ">Figure 7
<p>Illustration of the formation mechanism of the TA–PEG passivating coating on the pyrite surface in water-filled mining environments.</p>
Full article ">
47 pages, 6131 KiB  
Review
Introductory Review of Soft Implantable Bioelectronics Using Conductive and Functional Hydrogels and Hydrogel Nanocomposites
by San Kim, Yumin Shin, Jaewon Han, Hye Jin Kim and Sung-Hyuk Sunwoo
Gels 2024, 10(10), 614; https://doi.org/10.3390/gels10100614 - 25 Sep 2024
Viewed by 1790
Abstract
Interfaces between implantable bioelectrodes and tissues provide critical insights into the biological and pathological conditions of targeted organs, aiding diagnosis and treatment. While conventional bioelectronics, made from rigid materials like metals and silicon, have been essential for recording signals and delivering electric stimulation, [...] Read more.
Interfaces between implantable bioelectrodes and tissues provide critical insights into the biological and pathological conditions of targeted organs, aiding diagnosis and treatment. While conventional bioelectronics, made from rigid materials like metals and silicon, have been essential for recording signals and delivering electric stimulation, they face limitations due to the mechanical mismatch between rigid devices and soft tissues. Recently, focus has shifted toward soft conductive materials, such as conductive hydrogels and hydrogel nanocomposites, known for their tissue-like softness, biocompatibility, and potential for functionalization. This review introduces these materials and provides an overview of recent advances in soft hydrogel nanocomposites for implantable electronics. It covers material strategies for conductive hydrogels, including both intrinsically conductive hydrogels and hydrogel nanocomposites, and explores key functionalization techniques like biodegradation, bioadhesiveness, injectability, and self-healing. Practical applications of these materials in implantable electronics are also highlighted, showcasing their effectiveness in real-world scenarios. Finally, we discuss emerging technologies and future needs for chronically implantable bioelectronics, offering insights into the evolving landscape of this field. Full article
(This article belongs to the Special Issue Advances in Hydrogels and Hydrogel-Based Composites)
Show Figures

Figure 1

Figure 1
<p>Overall structure of soft implantable bioelectronics using hydrogel and hydrogel nanocomposites. (<b>a</b>) Mechanical modulus of biological tissue and representative materials constituting bioelectronics. (<b>b</b>,<b>c</b>) Microscopic image showing conformal contact of flexible plastic (<b>b</b>) and soft hydrogel (<b>c</b>) on the skin surface. (<b>d</b>) Schematic illustration describing overall structure of the review.</p>
Full article ">Figure 2
<p>Conductive hydrogel and hydrogel nanocomposite. (<b>a</b>) Flowchart showing the conducting mechanism of hydrogel. (<b>b</b>,<b>c</b>) Pure conductive hydrogels. (<b>b</b>) Reproduced with permission from Ref. [<a href="#B31-gels-10-00614" class="html-bibr">31</a>], Copyright 2019, Springer Nature. (<b>c</b>) Reproduced with permission from Ref. [<a href="#B32-gels-10-00614" class="html-bibr">32</a>], Copyright 2022, AAAS. (<b>d</b>,<b>e</b>) Hydrogels with ionic conduction. Reproduced with permission from Ref. [<a href="#B33-gels-10-00614" class="html-bibr">33</a>], Copyright 2022, Springer Nature. (<b>f</b>,<b>g</b>) Hydrogel nanocomposites with conductive carbon nanomaterials. Reproduced with permission from Ref. [<a href="#B34-gels-10-00614" class="html-bibr">34</a>], Copyright 2023, John Wiley &amp; Sons, Inc. (<b>h</b>,<b>i</b>) Hydrogel nanocomposites with conductive metallic nanofillers. (<b>h</b>) Reproduced with permission from Ref. [<a href="#B35-gels-10-00614" class="html-bibr">35</a>], Copyright 2014, American Chemical Society. (<b>i</b>) Reproduced with permission from Ref. [<a href="#B36-gels-10-00614" class="html-bibr">36</a>], Copyright 2019, AIP Publishing. (<b>j</b>,<b>k</b>) Conductive hydrogel nanocomposites containing liquid metals. (<b>j</b>) Reproduced with permission from Ref. [<a href="#B37-gels-10-00614" class="html-bibr">37</a>], Copyright 2024, Elsevier. (<b>k</b>) Reproduced with permission from Ref. [<a href="#B38-gels-10-00614" class="html-bibr">38</a>], Copyright 2021, Springer Nature.</p>
Full article ">Figure 3
<p>Advanced properties of soft implantable hydrogels. (<b>a</b>–<b>c</b>) Biodegradable hydrogels. (<b>a</b>) Reproduced with permission from Ref. [<a href="#B130-gels-10-00614" class="html-bibr">130</a>], Copyright 2022, American Chemical Society. (<b>b</b>) Reproduced with permission from Ref. [<a href="#B131-gels-10-00614" class="html-bibr">131</a>], Copyright 2018, American Chemical Society. (<b>c</b>) Reproduced with permission from Ref. [<a href="#B132-gels-10-00614" class="html-bibr">132</a>], Copyright 2024, American Chemical Society. (<b>d</b>–<b>f</b>) Tissue-adhesive hydrogels. (<b>d</b>) Reproduced with permission from Ref. [<a href="#B133-gels-10-00614" class="html-bibr">133</a>], Copyright 2022, AAAS. (<b>e</b>) Reproduced with permission from Ref. [<a href="#B134-gels-10-00614" class="html-bibr">134</a>], Copyright 2023, Elsevier. (<b>f</b>) Reproduced with permission from Ref. [<a href="#B135-gels-10-00614" class="html-bibr">135</a>], Copyright 2018, American Chemical Society. (<b>g</b>–<b>i</b>) Injectable hydrogels. (<b>g</b>) Reproduced with permission from Ref. [<a href="#B136-gels-10-00614" class="html-bibr">136</a>], Copyright 2022, John Wiley &amp; Sons, Inc. (<b>h</b>) Reproduced with permission from Ref. [<a href="#B137-gels-10-00614" class="html-bibr">137</a>], Copyright 2023, John Wiley &amp; Sons, Inc. (<b>i</b>) Reproduced with permission from Ref. [<a href="#B138-gels-10-00614" class="html-bibr">138</a>], Copyright 2021, AAAS. (<b>j</b>–<b>l</b>) Self-healing hydrogels. (<b>j</b>) Reproduced with permission from Ref. [<a href="#B139-gels-10-00614" class="html-bibr">139</a>], Copyright 2017, Springer Nature. (<b>k</b>) Reproduced with permission from Ref. [<a href="#B140-gels-10-00614" class="html-bibr">140</a>], Copyright 2023, Springer Nature. (<b>l</b>) Reproduced with permission from Ref. [<a href="#B139-gels-10-00614" class="html-bibr">139</a>], Copyright 2017, Springer Nature.</p>
Full article ">Figure 4
<p>Functional conductive hydrogels for monitoring and therapeutic applications, classified into: (<b>a</b>–<b>c</b>) Monitoring biological signals; (<b>d</b>–<b>f</b>) therapeutic applications. (<b>a</b>) Brain activities monitoring. (<b>b</b>) Cardiovascular system monitoring. (<b>c</b>) Other types of biological signal monitoring. (<b>d</b>) MI therapy. (<b>e</b>) Tissue regeneration. (<b>f</b>) Wound healing.</p>
Full article ">Figure 5
<p>Conductive hydrogels and nanocomposites in soft implantable electronics for brain neural recording. (<b>a</b>) PEDOT:PSS conductive hydrogel electrode fabricated by laser-induced phase separation. Reproduced with permission from Ref. [<a href="#B32-gels-10-00614" class="html-bibr">32</a>], Copyright 2022, AAAS. (<b>b</b>) Injectable and tissue-conformable conductive hydrogel for MRI-compatible brain electrodes. Reproduced with permission from Ref. [<a href="#B307-gels-10-00614" class="html-bibr">307</a>], Copyright 2023, OAE Publishing, Inc. (<b>c</b>) 3D printable conducting polymer for soft neural probes. Reproduced with permission from Ref. [<a href="#B308-gels-10-00614" class="html-bibr">308</a>], Copyright 2020, Springer Nature. (<b>d</b>) Fully viscoelastic electrode for stimulation and recording of electrocorticograms on a rat cortical surface. Reproduced with permission from Ref. [<a href="#B309-gels-10-00614" class="html-bibr">309</a>], Copyright 2021, Springer Nature. (<b>e</b>) Hydrogel optoelectronic device (optrode) for electrophysiology recording signals from the mouse VTA. Reproduced with permission from Ref. [<a href="#B310-gels-10-00614" class="html-bibr">310</a>], Copyright 2024, Springer Nature.</p>
Full article ">Figure 6
<p>Conductive hydrogels and nanocomposites in soft implantable electronics for ECG monitoring and MI therapy. (<b>a</b>) Chronological adhesive hydrogel patch for synchronous mechanophysiological monitoring and electrocoupling therapy. Reproduced with permission from Ref. [<a href="#B300-gels-10-00614" class="html-bibr">300</a>], Copyright 2023, Springer Nature. (<b>b</b>) Highly conductive tissue-like hydrogel interface based on template-directed in-situ synthesis. Reproduced with permission from Ref. [<a href="#B311-gels-10-00614" class="html-bibr">311</a>], Copyright 2023, Springer Nature. (<b>c</b>) Fully viscoelastic electrode on a mouse heart, showing the recorded ECG with filtered and averaged beats. Reproduced with permission from Ref. [<a href="#B309-gels-10-00614" class="html-bibr">309</a>], Copyright 2021, Springer Nature. (<b>d</b>) 3D-printed hydrogel bioelectronics for electrophysiological monitoring and electrical modulation. Reproduced with permission from Ref. [<a href="#B312-gels-10-00614" class="html-bibr">312</a>], Copyright 2023, John Wiley and Sons. (<b>e</b>) 3D-printable high-performance conducting polymer hydrogel for all-hydrogel bioelectronic interfaces. Reproduced with permission from Ref. [<a href="#B30-gels-10-00614" class="html-bibr">30</a>], Copyright 2023, Springer Nature.</p>
Full article ">Figure 7
<p>Conductive hydrogels and nanocomposites in soft implantable electronics for in vivo neural signal recording and stimulation in the nervous system. (<b>a</b>) A 3D-printable high-performance conducting polymer hydrogel for all-hydrogel bioelectronic interfaces. Reproduced with permission from Ref. [<a href="#B301-gels-10-00614" class="html-bibr">301</a>], Copyright 2023, Springer Nature. (<b>b</b>) Highly conductive tissue-like hydrogel interface through template-directed assembly. Reproduced with permission from Ref. [<a href="#B293-gels-10-00614" class="html-bibr">293</a>], Copyright 2023, Springer Nature. (<b>c</b>) Highly stable, injectable, conductive hydrogel for chronic neuromodulation. Reproduced with permission from Ref. [<a href="#B313-gels-10-00614" class="html-bibr">313</a>], Copyright 2024, Springer Nature. (<b>d</b>) Injectable IT-IC hydrogel interfacing for instantaneous closed-loop rehabilitation. Reproduced with permission from Ref. [<a href="#B300-gels-10-00614" class="html-bibr">300</a>], Copyright 2023, Springer Nature. <span class="html-italic">p</span>-values are indicated with asterisks: * <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 ">
19 pages, 5022 KiB  
Article
Anti-Biofouling Polyzwitterion–Poly(amidoxime) Composite Hydrogel for Highly Enhanced Uranium Extraction from Seawater
by Lang Yang, Ye Sun, Yue Sun, Jiawen Wang, Lin Chen, Xueliang Feng, Jinggang Wang, Ning Wang, Dong Zhang and Chunxin Ma
Gels 2024, 10(9), 603; https://doi.org/10.3390/gels10090603 - 22 Sep 2024
Viewed by 994
Abstract
Amidoxime-functionalized hydrogels are one of most promising adsorbents for high-efficiency uranium (U) extraction from seawater, but bioadhesion on their surface seriously decreases their adsorption efficiency and largely shortens their service life. Herein, a semi-interpenetrating zwitterion–poly(amidoxime) (ZW-PAO) hydrogel was explored through introducing a PAO [...] Read more.
Amidoxime-functionalized hydrogels are one of most promising adsorbents for high-efficiency uranium (U) extraction from seawater, but bioadhesion on their surface seriously decreases their adsorption efficiency and largely shortens their service life. Herein, a semi-interpenetrating zwitterion–poly(amidoxime) (ZW-PAO) hydrogel was explored through introducing a PAO polymer into a poly [3-(dimethyl 4-vinylbenzyl amino) propyl sulfonate] (PDVBAP) polyzwitterionic (PZW) network via ultraviolet (UV) polymerization. Owing to the anti-polyelectrolyte effect of the PZW network, this ZW-PAO hydrogel can provide excellent super-hydrophilicity in seawater for high-efficiency U-adsorption from seawater. Furthermore, the ZW-PAO hydrogel had outstanding anti-biofouling performance for both highly enhanced U-adsorption and a relatively long working life in natural seawater. As a result, during only 25 days in seawater (without filtering bacteria), the U-uptake amount of this ZW-PAO hydrogel can reach 9.38 mg/g and its average rate can reach 0.375 mg/(g∙day), which is excellent among reported adsorbents. This work has explored a promising hydrogel for high-efficiency U-recovery from natural seawater and will inspire new strategy for U-adsorbing materials. Full article
(This article belongs to the Special Issue Advances in Functional and Intelligent Hydrogels)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The FT-IR spectra of the ZW-PAO hydrogel, PZW hydrogel, PAO, and PAN; (<b>b</b>) the EDS mappings of the ZW-PAO hydrogel and PAM-PAO hydrogel.</p>
Full article ">Figure 2
<p>(<b>a</b>) The photos of the switching of the ZW-PAO hydrogel and PAM-PAO hydrogel in seawater and pure water; (<b>b</b>) swelling ratio of the ZW-PAO hydrogel in pure water and seawater water.</p>
Full article ">Figure 3
<p>The characterization of the ZW-PAO hydrogel before and after U-extraction: (<b>a</b>) the XPS spectra; (<b>b</b>) the SEM and EDS mapping images.</p>
Full article ">Figure 4
<p>The U-adsorption kinetics of the ZW-PAO hydrogel and the pseudo-second-order in 2, 4, 8, and 16 ppm U-added (<b>a</b>,<b>b</b>) pure water and (<b>c</b>,<b>d</b>) seawater. All the solutions were adjusted to pH = 6 prior to U-adsorption.</p>
Full article ">Figure 5
<p>(<b>a</b>) U-extraction capacity as a function of pH in pure water containing 8 ppm U; (<b>b</b>) selective adsorption of metal ions by the ZW-PAO hydrogel.</p>
Full article ">Figure 6
<p>(<b>a</b>) U-adsorbing–desorbing process, (<b>b</b>) U-desorption kinetic curve and (<b>c</b>) 5 adsorbing–desorbing cycles of the ZW-PAO hydrogel.</p>
Full article ">Figure 7
<p>(<b>a</b>–<b>c</b>) The confocal laser microscopy of different ZW-PAO and PAM-PAO hydrogels; (<b>d</b>) the antibacterial rates of the ZW-PAO hydrogel against three types of bacteria; (<b>e</b>) the enhanced U-adsorption capacity of the ZW-PAO hydrogel by the improved anti-biofouling in seawater.</p>
Full article ">Figure 8
<p>(<b>a</b>) Comparison of real seawater adsorption kinetics between ZW-PAO and PAM-PAO hydrogels; (<b>b</b>) comparison of uranium extraction rate between ZW-PAO hydrogel and existing amidoxime-group adsorbents [<a href="#B20-gels-10-00603" class="html-bibr">20</a>,<a href="#B21-gels-10-00603" class="html-bibr">21</a>,<a href="#B39-gels-10-00603" class="html-bibr">39</a>,<a href="#B40-gels-10-00603" class="html-bibr">40</a>,<a href="#B41-gels-10-00603" class="html-bibr">41</a>,<a href="#B42-gels-10-00603" class="html-bibr">42</a>,<a href="#B43-gels-10-00603" class="html-bibr">43</a>,<a href="#B44-gels-10-00603" class="html-bibr">44</a>,<a href="#B45-gels-10-00603" class="html-bibr">45</a>,<a href="#B46-gels-10-00603" class="html-bibr">46</a>,<a href="#B47-gels-10-00603" class="html-bibr">47</a>,<a href="#B48-gels-10-00603" class="html-bibr">48</a>,<a href="#B49-gels-10-00603" class="html-bibr">49</a>,<a href="#B50-gels-10-00603" class="html-bibr">50</a>,<a href="#B51-gels-10-00603" class="html-bibr">51</a>,<a href="#B52-gels-10-00603" class="html-bibr">52</a>].</p>
Full article ">Scheme 1
<p>The schematic process of (<b>a</b>) the preparation of the ZW-PAO hydrogel and (<b>b</b>) selective U-adsorption by the ZW-PAO hydrogel.</p>
Full article ">
23 pages, 4533 KiB  
Article
Exploring Cationic Guar Gum: Innovative Hydrogels and Films for Enhanced Wound Healing
by Kamila Gabrieli Dallabrida, Willer Cezar Braz, Crisleine Marchiori, Thainá Mayer Alves, Luiza Stolz Cruz, Giovanna Araujo de Morais Trindade, Patrícia Machado, Lucas Saldanha da Rosa, Najeh Maissar Khalil, Fabiane Gomes de Moraes Rego, André Ricardo Fajardo, Luana Mota Ferreira, Marcel Henrique Marcondes Sari and Jéssica Brandão Reolon
Pharmaceutics 2024, 16(9), 1233; https://doi.org/10.3390/pharmaceutics16091233 - 22 Sep 2024
Cited by 1 | Viewed by 1279
Abstract
Background/Objectives: This study developed and characterized hydrogels (HG-CGG) and films (F-CGG) based on cationic guar gum (CGG) for application in wound healing. Methods: HG-CGG (2% w/v) was prepared by gum thickening and evaluated for pH, stability, spreadability, and viscosity. F-CGG [...] Read more.
Background/Objectives: This study developed and characterized hydrogels (HG-CGG) and films (F-CGG) based on cationic guar gum (CGG) for application in wound healing. Methods: HG-CGG (2% w/v) was prepared by gum thickening and evaluated for pH, stability, spreadability, and viscosity. F-CGG was obtained using an aqueous dispersion of CGG (6% w/v) and the solvent casting method. F-CGG was characterized for thickness, weight uniformity, morphology, mechanical properties, hydrophilicity, and swelling potential. Both formulations were evaluated for bioadhesive potential on intact and injured porcine skin, as well as antioxidant activity. F-CGG was further studied for biocompatibility using hemolysis and cell viability assays (L929 fibroblasts), and its wound-healing potential by the scratch assay. Results: HG-CGG showed adequate viscosity and spreadability profiles for wound coverage, but its bioadhesive strength was reduced on injured skin. In contrast, F-CGG maintained consistent bioadhesive strength regardless of skin condition (6554.14 ± 540.57 dyne/cm2 on injured skin), presenting appropriate mechanical properties (flexible, transparent, thin, and resistant) and a high swelling capacity (2032 ± 211% after 6 h). F-CGG demonstrated superior antioxidant potential compared to HG-CGG (20.50 mg/mL ABTS+ radical scavenging activity), in addition to exhibiting low hemolytic potential and no cytotoxicity to fibroblasts. F-CGG promoted the proliferation of L929 cells in vitro, supporting wound healing. Conclusions: Therefore, CGG proved to be a promising material for developing formulations with properties suitable for cutaneous use. F-CGG combines bioadhesion, antioxidant activity, biocompatibility, cell proliferation, and potential wound healing, making it promising for advanced wound treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of cationic guar gum (CGG).</p>
Full article ">Figure 2
<p>Macroscopic appearance of hydrogels during development formulation tests at different concentrations of CGG. Hydrogels at 1% (<b>A</b>); 2% (<b>B</b>); 3% (<b>C</b>); and 6% (<b>D</b>) <span class="html-italic">w</span>/<span class="html-italic">w</span> of CGG.</p>
Full article ">Figure 3
<p>HG-CGG spreadability profile (<b>A</b>) and viscogram (<b>B</b>).</p>
Full article ">Figure 4
<p>Images of films prepared with 1% (<b>A</b>), 2% (<b>B</b>), 4% (<b>C</b>), and 6% (<b>D</b>) (<span class="html-italic">w</span>/<span class="html-italic">v</span>) CGG during development tests and scans obtained in UV-Vis spectrum to F-CGG (6% <span class="html-italic">w</span>/<span class="html-italic">v</span>) (<b>E</b>).</p>
Full article ">Figure 5
<p>Scanning electron microscopy (SEM) images obtained from the surface portion of the F-CGG film using different magnifications (<b>A</b>–<b>C</b>) and SEM images obtained from the transversal section of F-CGG after cryofracture (<b>D</b>) (scale bars: (<b>A</b>) 500 µm, (<b>B</b>) 200 µm, (<b>C</b>) 100 µm, and (<b>D</b>) 100 µm).</p>
Full article ">Figure 6
<p>Swelling index (<b>A</b>) and representative images of contact angle (<b>B</b>) determination evaluated for F-CGG (<span class="html-italic">n</span> = 3). The results are expressed as mean ± SD (<span class="html-italic">n</span> = 3). Unpaired <span class="html-italic">t</span>-test. <span class="html-italic">p</span> &lt; 0.001 (***): significant difference between the contact angle obtained on the upper and lower faces of the F-CGG.</p>
Full article ">Figure 7
<p>ATR-FTIR of the pure CGG (<b>A</b>), F-CGG (<b>B</b>), and HG (<b>C</b>) and a comparison of them (<b>D</b>).</p>
Full article ">Figure 8
<p>PCA model applied to the ATR-FTIR spectra. (<b>A</b>) Eigenvalues against the number of principal components, and (<b>B</b>) score plot.</p>
Full article ">Figure 9
<p>Bioadhesive strength of HG-CGG and F-CGG in uninjured and injured skin. Mean ± SD (<span class="html-italic">n</span> = 3). One-way ANOVA followed by Tukey’s test: (***) difference between HG-CGG in uninjured and injured skin (<span class="html-italic">p</span> &lt; 0.001); (**) difference between HG-CGG and F-CGG in both uninjured and injured skin (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 10
<p>Antioxidant activity of HG-CGG and F-CGG. Mean ± SD (<span class="html-italic">n</span> = 3). One-way ANOVA followed by Tukey’s test: (*) difference between HG-CGG and F-CGG in 5.13 mg/mL (<span class="html-italic">p</span> &lt; 0.05); (***) difference between HG-CGG and F-CGG in 10.25 and 20.50 mg/mL (<span class="html-italic">p</span> &lt; 0.001). For this assay, formulations were weighed (2.56 to 20.5 mg/mL of the formulations), corresponding to a CGG concentration range of 0.055 to 0.44 mg/mL to HG and 1.40 to 11.22 mg/mL to film.</p>
Full article ">Figure 11
<p>Hemolytic effect (<b>A</b>) and cell viability of L-929 cells after 24 h of treatment with F-CGG (<b>B</b>). Mean ± SD (<span class="html-italic">n</span> = 3). One-way ANOVA followed by Tukey’s test: (***) difference compared to the 20.5 mg group. For the assay in (<b>A</b>), film fragments were weighed (20.5, 41, and 82 mg of formulation/tube), corresponding to the CGG concentration ranges of 11.22, 22.44, and 44.88 mg/tube. In (<b>B</b>), the range of 50–1000 µg/mL of the formulation corresponded to 27–547 µg/mL of gum.</p>
Full article ">Figure 12
<p>Wound-healing properties of 50 and 1000 μg/mL on L-929 cells (fibroblasts) at 0, 4, 8, 12, and 24 h after the treatment (T; time). DMEM was used as a control. Pictures were taken in the phase contrast mode, 10×, scale bar = 200 μm.</p>
Full article ">
15 pages, 2799 KiB  
Article
Bio-Epoxy Resins Based on Lignin and Tannic Acids as Wood Adhesives—Characterization and Bonding Properties
by Ivana Gavrilović-Grmuša, Milica Rančić, Tamara Tešić, Stevan Stupar, Milena Milošević and Jelena Gržetić
Polymers 2024, 16(18), 2602; https://doi.org/10.3390/polym16182602 - 14 Sep 2024
Viewed by 1383
Abstract
The possibility of producing and designing bio-epoxides based on the natural polyphenol lignin/epoxidized lignin and tannic acids for application as wood adhesives is presented in this work. Lignin and tannic acids contain numerous reactive hydroxyl phenolic moieties capable of being efficiently involved in [...] Read more.
The possibility of producing and designing bio-epoxides based on the natural polyphenol lignin/epoxidized lignin and tannic acids for application as wood adhesives is presented in this work. Lignin and tannic acids contain numerous reactive hydroxyl phenolic moieties capable of being efficiently involved in the reaction with commercial epoxy resins as a substitute for commercial, non-environmentally friendly, toxic amine-based hardeners. Furthermore, lignin was epoxidized in order to obtain an epoxy lignin that can be a replacement for diglycidyl ether bisphenol A (DGEBA). Cross-linking of bio-epoxy epoxides was investigated via FTIR spectroscopy and their prospects for wood adhesive application were evaluated. This study determined that the curing reaction of epoxy resin can be conducted using lignin/epoxy lignin or tannic acid. Tensile shear strength testing results showed that lignin and tannic acid can effectively replace amine hardeners in epoxy resins. Examination of the failure of the samples showed that all samples had a 100% fracture through the wood. All samples of bio-epoxy adhesives displayed significant tensile shear strength in the range of 5.84–10.87 MPa. This study presents an innovative approach to creating novel cross-linked networks of eco-friendly and high-performance wood bio-adhesives. Full article
(This article belongs to the Special Issue Recent Developments in Biodegradable and Biobased Polymers II)
Show Figures

Figure 1

Figure 1
<p>Cross-linking reactions of tannic acids (TA), Kraft lignin (AKL), and epoxy-modified lignin (EL) with ERs.</p>
Full article ">Figure 2
<p>Mechanism of lignin epoxidation reaction by epichlorohydrin in alkaline conditions to obtain epoxy lignin (EL).</p>
Full article ">Figure 3
<p>FTIR spectra of the (<b>a</b>) normalized FTIR (<b>b</b>) raw materials and (<b>c</b>) cured control epoxy and TA/L bio-epoxy adhesives.</p>
Full article ">Figure 4
<p>Mechanism of the crosslinking of DGEBA, epoxy lignin, and amino hardener during curing.</p>
Full article ">Figure 5
<p>Mechanism of the crosslinking of epoxy lignin and tannic acid during curing.</p>
Full article ">Figure 6
<p>(<b>a</b>) preparing the samples for tensile shear strength determination; (<b>b</b>) samples prepared and conditioned for tensile shear strength determination; (<b>c</b>) joint sample after tensile shear testing that have undergone a 100% fracture through the wood.</p>
Full article ">Figure 7
<p>Tensile shear strength results for different adhesive blends.</p>
Full article ">Figure 8
<p>(<b>a</b>) ER-AH120; (<b>b</b>) ER-TA10; (<b>c</b>) ER-TA-EL10; and (<b>d</b>) ER-TA-L10.</p>
Full article ">
Back to TopTop