[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
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (497)

Search Parameters:
Keywords = sheared gel

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
31 pages, 4890 KiB  
Article
Characteristics of Hydrogels as a Coating for Microneedle Transdermal Delivery Systems with Agomelatine
by Monika Wojtyłko, Ariadna B. Nowicka, Anna Froelich, Mirosław Szybowicz, Tobiasz Banaszek, Dorota Tomczak, Wiesław Kuczko, Radosław Wichniarek, Irena Budnik, Barbara Jadach, Oliwia Kordyl, Antoni Białek, Julia Krysztofiak, Tomasz Osmałek and Dimitrios A. Lamprou
Molecules 2025, 30(2), 322; https://doi.org/10.3390/molecules30020322 - 15 Jan 2025
Viewed by 674
Abstract
Agomelatine (AGM) is an effective antidepressant with low oral bioavailability due to intensive hepatic metabolism. Transdermal administration of agomelatine may increase its bioavailability and reduce the doses necessary for therapeutic effects. However, transdermal delivery requires crossing the stratum corneum barrier. For this purpose, [...] Read more.
Agomelatine (AGM) is an effective antidepressant with low oral bioavailability due to intensive hepatic metabolism. Transdermal administration of agomelatine may increase its bioavailability and reduce the doses necessary for therapeutic effects. However, transdermal delivery requires crossing the stratum corneum barrier. For this purpose, the use of microneedles may increase the efficiency of administration. The aim of this study was to prepare an agomelatine-loaded hydrogel suitable for coating microneedles for the transdermal drug delivery of AGM. The optimized formulations were subjected to spectroscopic and rheological characterization and mechanical tests, as well as tested for release through an artificial membrane and permeation through human skin ex vivo. Both hydrogels were found to have suitable parameters for coating microneedles using the dip-coating method, including the stability of the substance at the process temperature, shear-thinning behavior, and appropriate textural parameters such as adhesion or hardness. Additionally, two formulations were tested for potential application to the skin alone because the gels showed suitable mechanical properties for the skin application. In this case, the ethanol gel was characterized by higher skin permeability and better spreadability. The information obtained in this study will allow the preparation of coated microneedles for the transdermal administration of agomelatine. Full article
(This article belongs to the Special Issue Hydrogels: Preparation, Characterization, and Applications)
Show Figures

Figure 1

Figure 1
<p>Images of the formulations obtained by the optical microscope Targano Prestige: Sus placebo (<b>a</b>), Sus AGM (<b>b</b>,<b>c</b>), Et placebo (<b>d</b>), Et AGM (<b>e</b>,<b>f</b>); objective: 10× (<b>a</b>,<b>b</b>,<b>d</b>,<b>e</b>) and 25× (<b>c</b>,<b>f</b>).</p>
Full article ">Figure 2
<p>Length and width measurement results presented as the frequency of values in size ranges. “(”—range opened, “]”—range closed.</p>
Full article ">Figure 3
<p>The microscopic photo of AGM crystals present in the Sus AGM gel sample. The presented image was captured at a magnification of 400×.</p>
Full article ">Figure 4
<p>Calculation (DFT) and experimental Raman scattering spectra of agomelatine II form powders.</p>
Full article ">Figure 5
<p>Temperature dependences of AGM powder.</p>
Full article ">Figure 6
<p>The FWHM analysis at the band at 1370 cm<sup>−1</sup> (<b>top</b>). Comparison with the DSC results (<b>bottom</b>).</p>
Full article ">Figure 7
<p>Raman spectra of AGM and AGM in ethanol solution.</p>
Full article ">Figure 8
<p>Raman spectra of AGM, components of gels (glycerol, Carbopol, triisopropanolamine), and formulations with and without AGM.</p>
Full article ">Figure 9
<p>DSC thermograms of AGM II (black), Sus AGM (red), Sus placebo (green), Et AGM (blue), and Et placebo (cyan).</p>
Full article ">Figure 10
<p>Loss of volatile components during 210 min of drying.</p>
Full article ">Figure 11
<p>Texture profiles of placebo gels and gels loaded with agomelatine.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) Comparison of the values from texture profile analysis for the ethanol gel and suspension, both with the placebo and loaded with agomelatine. Adhesiveness is shown as an absolute value.</p>
Full article ">Figure 13
<p>Spreadability study of placebo gels and gels loaded with agomelatine.</p>
Full article ">Figure 14
<p>(<b>a</b>–<b>d</b>) Comparison of the results from the spreadability study for the ethanol gel and suspension, both with the placebo and loaded with agomelatine. Work of adhesion and force of adhesion are shown as absolute values.</p>
Full article ">Figure 15
<p>(<b>a</b>,<b>b</b>) Flow curves for the placebo and agomelatine-loaded formulations obtained at 25 °C (<b>a</b>) and 32 °C (<b>b</b>).</p>
Full article ">Figure 16
<p>(<b>a</b>,<b>b</b>) Cross point of storage modulus G′ and loss modulus G″ at 25 °C (<b>a</b>) and 32 °C (<b>b</b>).</p>
Full article ">Figure 17
<p>(<b>a</b>,<b>b</b>) Storage modulus G′ and loss modulus G″ in a variable frequency of oscillation at 25 °C (<b>a</b>) and 32 °C (<b>b</b>).</p>
Full article ">Figure 18
<p>Cumulative amount of drug released through the membrane per surface area [µg/cm<sup>2</sup>].</p>
Full article ">Figure 19
<p>Comparison of the linear parts of the graphs. Up to 720 min for Sus AGM and a part up to 60 min for Et AGM (extrapolated for better visibility) [µg/cm<sup>2</sup>].</p>
Full article ">Figure 20
<p>(<b>a</b>,<b>b</b>) Cumulative amount of drug permeated through the skin per surface area [µg/cm<sup>2</sup>].</p>
Full article ">Figure 21
<p>Amount of drug retained in different layers of the skin.</p>
Full article ">
20 pages, 2844 KiB  
Article
Rheology and Stability of Hydrocarbon-Based Gelled Fuels for Airbreathing Applications
by Simone Dell’Acqua, Francesco Morando, Stefania Carlotti and Filippo Maggi
Aerospace 2025, 12(1), 49; https://doi.org/10.3390/aerospace12010049 - 13 Jan 2025
Viewed by 416
Abstract
Gelled fuels are rheologically complex, non-Newtonian fluids. They combine the benefits of both liquid and solid states, reducing risks of leakage, spilling, and sloshing during storage while maintaining the ability to be sprayed inside a combustion chamber. Additionally, suspending energetic particles, such as [...] Read more.
Gelled fuels are rheologically complex, non-Newtonian fluids. They combine the benefits of both liquid and solid states, reducing risks of leakage, spilling, and sloshing during storage while maintaining the ability to be sprayed inside a combustion chamber. Additionally, suspending energetic particles, such as metal powders of aluminum and boron, can significantly enhance their energy density compared to conventional liquid fuels. In this study, several kerosene-based and ethanol-based formulations were experimentally investigated, using both organic and inorganic gelling agents. The compositions were optimized in terms of the gellant amount and manufacturing process. Some of the most promising gellants for kerosene include fatty acids, such as Thixcin® R or THIXATROL® ST, and metallic soaps, such as aluminum stearate and zinc stearate. The effects of various co-solvents were assessed, including ketones (methyl isoamyl ketone, methyl ethyl ketone, and acetone) and alcohols (ethanol and octadecanol). Sugar polymers like hydroxypropyl cellulose were tested as gelling agents for ethanol. A preliminary rheological analysis was conducted to characterize their behavior at rest and under shear stress. Finally, a novel approach was introduced to study the stability of the gels under vibration, which was derived from a realistic mission profile of a ramjet. Finally, the ideal gravimetric specific impulse was evaluated through ideal thermochemical computations. The results showed that promising formulations can be found in both kerosene-based and ethanol-based gels. Such compositions are of interest in practical airbreathing applications as they have demonstrated excellent stability under vibration, ideal combustion properties, and pronounced shear-thinning behavior. Full article
Show Figures

Figure 1

Figure 1
<p>Conceptual scheme of the experimental setup implemented for the vibration assessment of gelled fuels.</p>
Full article ">Figure 2
<p>Picture of the experimental setup for the vibration tests. The power amplifier is not shown in the figure.</p>
Full article ">Figure 3
<p>Qualitative representation of gel liquefaction behavior under vibration, describing the best (<b>a</b>) and worst (<b>b</b>) performances. Grade A2 is associated with an intermediate performance.</p>
Full article ">Figure 4
<p>Kerosene–aluminum stearate gels, highlighting the appearance in different preparation phases. Immediately after the original preparation (<b>a</b>); after a few hours (<b>b</b>); after the second mixing (<b>c</b>).</p>
Full article ">Figure 5
<p>Thermogravimetry analysis results for the most promising gels (<b>a</b>) and for gel components (<b>b</b>).</p>
Full article ">Figure 6
<p>Frequency-sweep dynamic measurements for different strain percentages.</p>
Full article ">Figure 7
<p>Dynamic strain–sweep analysis outcomes. (<b>a</b>) Storage modulus–shear stress curves, from which the yield stress <math display="inline"><semantics> <msub> <mi>τ</mi> <mrow> <mi>y</mi> <mi>i</mi> <mi>e</mi> <mi>l</mi> <mi>d</mi> </mrow> </msub> </semantics></math> can be derived; (<b>b</b>) storage modulus (black) and loss modulus (gray) profiles, identifying the flow point <math display="inline"><semantics> <msub> <mi>γ</mi> <mrow> <mi>f</mi> <mi>l</mi> <mi>o</mi> <mi>w</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 8
<p>Rheological measurements from steady tests represented in logarithmic axes.</p>
Full article ">Figure 9
<p>Gravimetric specific impulse for gelled propellants and pure liquid fuels as a function of the air-to-fuel mass ratio.</p>
Full article ">
25 pages, 6486 KiB  
Article
Thermoresponsive Gels with Rosemary Essential Oil: A Novel Topical Carrier for Antimicrobial Therapy and Drug Delivery Applications
by Ludovic Everard Bejenaru, Adina-Elena Segneanu, Cornelia Bejenaru, Ionela Amalia Bradu, Titus Vlase, Dumitru-Daniel Herea, Marius Ciprian Văruţ, Roxana Maria Bălăşoiu, Andrei Biţă, Antonia Radu, George Dan Mogoşanu and Maria Viorica Ciocîlteu
Gels 2025, 11(1), 61; https://doi.org/10.3390/gels11010061 - 12 Jan 2025
Viewed by 567
Abstract
This study investigates the development and comprehensive characterization of innovative thermoresponsive gels incorporating rosemary essential oil (RoEO) encapsulated in poly(lactic-co-glycolic acid) (PLGA) microparticles, with a focus on their potential applications in topical antimicrobial and wound healing therapies. RoEO, renowned for its [...] Read more.
This study investigates the development and comprehensive characterization of innovative thermoresponsive gels incorporating rosemary essential oil (RoEO) encapsulated in poly(lactic-co-glycolic acid) (PLGA) microparticles, with a focus on their potential applications in topical antimicrobial and wound healing therapies. RoEO, renowned for its robust antimicrobial, antioxidant, and wound-healing properties, was subjected to detailed chemical profiling using gas chromatography-mass spectrometry (GC–MS), which identified oxygenated monoterpenes as its dominant constituents. PLGA microparticles were synthesized through an optimized oil-in-water emulsion technique, ensuring high encapsulation efficiency and structural integrity. These microparticles were thoroughly characterized using Fourier-transform infrared (FTIR) spectroscopy to confirm functional group interactions, scanning electron microscopy (SEM) for surface morphology, X-ray diffraction (XRD) for crystalline properties, and thermal analysis for stability assessment. The synthesized microparticles displayed uniform size distribution and efficient encapsulation, demonstrating compatibility with the gel matrix. Two distinct thermoresponsive gel formulations were developed using varying ratios of Poloxamer 407 and Poloxamer 188 to achieve optimal performance. The gels were evaluated for key physicochemical properties, including pH, gelation temperature, viscosity, and rheological behavior. Both formulations exhibited thermoresponsive gelation at skin-compatible temperatures (27.6 °C and 32.9 °C), favorable pH levels (6.63 and 6.40), and shear-thinning behavior suitable for topical application. Antimicrobial efficacy was assessed against common pathogens associated with skin infections, including Staphylococcus aureus, Escherichia coli, and Candida albicans. The RoEO-PLGA-loaded gels demonstrated significant inhibitory effects, confirming their potential as effective carriers for controlled and localized drug delivery. These findings underscore the promising application of RoEO-PLGA-loaded thermoresponsive gels in addressing challenges associated with topical antimicrobial therapies and wound care, offering an innovative approach to enhancing therapeutic outcomes. By integrating the bioactive potential of RoEO with the advanced delivery capabilities of PLGA microparticles and thermoresponsive gels, this study paves the way for the development of next-generation formulations tailored to meet the specific needs of localized drug delivery in skin health management. Full article
Show Figures

Figure 1

Figure 1
<p>GC–MS chromatogram of RoEO Tunisia reference. GC: Gas chromatography; MS: Mass spectrometry; RoEO: Rosemary essential oil.</p>
Full article ">Figure 2
<p>GC–MS chromatogram of RoEO sample.</p>
Full article ">Figure 3
<p>FTIR spectra for PLGA, RoEO, and RoEO-PLGA samples. FTIR: Fourier-transform infrared; PLGA: Poly(lactic-<span class="html-italic">co</span>-glycolic) acid; RoEO: Rosemary essential oil.</p>
Full article ">Figure 4
<p>XRD pattern of PLGA, RoEO, and RoEO-PLGA samples. XRD: X-ray diffraction.</p>
Full article ">Figure 5
<p>Morphological aspects of RoEO-PLGA microparticles (SEM image). SEM: Scanning electron microscopy.</p>
Full article ">Figure 6
<p>EDS spectrum of RoEO-PLGA microparticles. EDS: Energy-dispersive X-ray spectroscopy.</p>
Full article ">Figure 7
<p>DLS pattern of RoEO-PLGA sample.</p>
Full article ">Figure 8
<p>Thermoanalytical data for RoEO-PLGA sample (black line: TG analysis; green line: DTG analysis; red line: HF). DTG: Derivative thermogravimetry; HF: Heat flow; TG: Thermogravimetry.</p>
Full article ">Figure 9
<p>Rheological behavior of RoEO-PLGA _A: (<b>a</b>) Relationship between shear stress and shear rate (blue—increasing shear rate; red—decreasing shear rate); (<b>b</b>) Viscosity under varying shear rates (at 36 °C; blue—increasing shear rate; red—decreasing shear rate).</p>
Full article ">Figure 10
<p>Rheological behavior of RoEO-PLGA _B: (<b>a</b>) Relationship between shear stress and shear rate (blue—increasing shear rate; red—decreasing shear rate); (<b>b</b>) Viscosity under varying shear rates (at 36 °C; blue—increasing shear rate; red—decreasing shear rate).</p>
Full article ">Figure 11
<p>Antimicrobial screening against (<b>a</b>) <span class="html-italic">S. aureus</span>, (<b>b</b>) <span class="html-italic">E. coli</span>, and (<b>c</b>) a fungal strain (<span class="html-italic">C. albicans</span>).</p>
Full article ">Figure 12
<p>Graphic process of RoEO-PLGA microparticles and RoEO-PLGA gels formulation and characterization. DCM: Dichloromethane; DLS: Dynamic light scattering; FTIR: Fourier-transform infrared; PLGA: Poly(lactic-<span class="html-italic">co</span>-glycolic) acid; PVA: Poly(vinyl alcohol); RoEO: Rosemary essential oil; SEM: Scanning electron microscopy.</p>
Full article ">
17 pages, 4693 KiB  
Article
Rheological Characterization and Printability of Sodium Alginate–Gelatin Hydrogel for 3D Cultures and Bioprinting
by Mohan Kumar Dey and Ram V. Devireddy
Biomimetics 2025, 10(1), 28; https://doi.org/10.3390/biomimetics10010028 - 4 Jan 2025
Viewed by 643
Abstract
The development of biocompatible hydrogels for 3D bioprinting is essential for creating functional tissue models and advancing preclinical drug testing. This study investigates the formulation, printability, mechanical properties, and biocompatibility of a novel Alg-Gel hydrogel blend (alginate and gelatin) for use in extrusion-based [...] Read more.
The development of biocompatible hydrogels for 3D bioprinting is essential for creating functional tissue models and advancing preclinical drug testing. This study investigates the formulation, printability, mechanical properties, and biocompatibility of a novel Alg-Gel hydrogel blend (alginate and gelatin) for use in extrusion-based 3D bioprinting. A range of hydrogel compositions were evaluated for their rheological behavior, including shear-thinning properties, storage modulus, and compressive modulus, which are crucial for maintaining structural integrity during printing and supporting cell viability. The printability assessment of the 7% alginate–8% gelatin hydrogel demonstrated that the 27T tapered needle achieved the highest normalized Printability Index (POInormalized = 1), offering the narrowest strand width (0.56 ± 0.02 mm) and the highest printing accuracy (97.2%) at the lowest printing pressure (30 psi). In contrast, the 30R needle, with the smallest inner diameter (0.152 mm) and highest printing pressure (80 psi), resulted in the widest strand width (0.70 ± 0.01 mm) and the lowest accuracy (88.8%), resulting in a POInormalized of 0.274. The 30T and 27R needles demonstrated moderate performance, with POInormalized values of 0.758 and 0.558, respectively. The optimized 7% alginate and 8% gelatin blend demonstrated favorable printability, mechanical strength, and cell compatibility with MDA-MB-213 breast cancer cells, exhibiting high cell proliferation rates and minimal cytotoxicity over a 2-week culture period. This formulation offers a balanced approach, providing sufficient viscosity for precision printing while minimizing shear stress to preserve cell health. This work lays the groundwork for future advancements in bioprinted cancer models, contributing to the development of more effective tools for drug screening and personalized medicine. Full article
(This article belongs to the Section Biomimetic Design, Constructions and Devices)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Scaffold grid design with extruded square along the path line and (<b>b</b>) layered scaffold configuration with 90° rotation and z-axis duplication.</p>
Full article ">Figure 2
<p>(<b>a</b>) Laser-cut square-shaped molds and (<b>b</b>) casting process for Alg-Gel hydrogel samples.</p>
Full article ">Figure 3
<p>MDA-MB-231 cells seeded on 3D-bioprinted scaffolds in a 12-well plate.</p>
Full article ">Figure 4
<p>Rheological characterization of hydrogel mixtures with varying alginate and gelatin concentrations (4% Alg–8% Gel, 5% Alg–6% Gel, 5% Alg–6% Gel, 7% Alg–8% Gel). (<b>a</b>) Storage modulus (G′) and loss modulus (G″) as a function of angular frequency, showing an increase in both moduli with higher alginate concentration; (<b>b</b>) tan δ vs. angular frequency for the hydrogel mixtures, with tan δ values consistently below 1 across all formulations; (<b>c</b>) shear viscosity as a function of shear rate, demonstrating shear-thinning behavior in all hydrogel mixtures. (<b>d</b>) Axial stress vs. compression percentage, highlighting distinct mechanical behaviors across formulations, with 4% Alg–8% Gel showing the highest compressive strength.</p>
Full article ">Figure 5
<p>Swelling ratio of Alg-Gel hydrogels over time.</p>
Full article ">Figure 6
<p>UATR spectra of (<b>a</b>) alginate, (<b>b</b>) alginate–gelatin, and (<b>c</b>) alginate–gelatin–calcium chloride.</p>
Full article ">Figure 7
<p>Three-dimensional bioprinting scaffold on Petri dish.</p>
Full article ">Figure 8
<p>Evaluation of cell proliferation and viability of MDA-MB-213 cells cultured on Alg-Gel hydrogels over two weeks: (<b>a</b>) cell viability at 1 day; (<b>b</b>) cell viability at 1 week; (<b>c</b>) cell viability at 2 weeks, confirming hydrogel cytocompatibility and support for long-term culture; (<b>d</b>) cell viability.</p>
Full article ">
23 pages, 2532 KiB  
Article
Fabrication of Thymoquinone and Ascorbic Acid-Loaded Spanlastics Gel for Hyperpigmentation: In Vitro Release, Cytotoxicity, and Skin Permeation Studies
by Ahlam Zaid Alkilani, Rua’a Alkhaldi, Haneen A. Basheer, Bassam I. Amro and Maram A. Alhusban
Pharmaceutics 2025, 17(1), 48; https://doi.org/10.3390/pharmaceutics17010048 - 2 Jan 2025
Viewed by 645
Abstract
Background/Objectives: The demand for a safe compound for hyperpigmentation is continuously increasing. Bioactive compounds such as thymoquinone (TQ) and ascorbic acid (AA) induce inhibition of melanogenesis with a high safety profile. The aim of this study was to design and evaluate spanlastics [...] Read more.
Background/Objectives: The demand for a safe compound for hyperpigmentation is continuously increasing. Bioactive compounds such as thymoquinone (TQ) and ascorbic acid (AA) induce inhibition of melanogenesis with a high safety profile. The aim of this study was to design and evaluate spanlastics gel loaded with bioactive agents, TQ and AA, for the management of hyperpigmentation. Methods: Several spanlastics formulations were successfully fabricated and characterized in terms of morphology, vesicle size, zeta potential, and release. Results: The optimized TQ-loaded spanlastic formulation showed an average size of 223.40 ± 3.50 nm, and 133.00 ± 2.80 nm for AA-loaded spanlastic formulation. The optimized spanlastics formulation showed the highest entrapment efficiency (EE%) of 97.18 ± 2.02% and 93.08 ± 1.95%, for TQ and AA, respectively. Additionally, the edge activator concentration had a significant effect (p < 0.05) on EE%; it was found that by increasing the amount of EA, the EE% increases. Following that, the optimal spanlastics fomulation loaded with TQ and AA were incorporated into gel and explored for appearance, pH, spreadability, stability, rheology, in vitro release, ex vivo permeation study, and MTT cytotoxicity. The formulated spanlastics gel (R-1) has a pH of 5.53. Additionally, R-1 gel was significantly (p < 0.05) more spreadable than control gel, and exhibited a shear thinning behavior. Most importantly, ex vivo skin deposition studies confirmed superior skin deposition of TQ and AA from spanlastic gels. Additionally, results indicated that tyrosinase inhibition was primarily due to TQ. When comparing TQ alone with the TQ-AA combination, inhibition ranged from 18.35 to 42.73% and 24.28 to 42.53%, respectively. Both TQ spanlastics and the TQ-AA combination showed a concentration-dependent inhibition of tyrosinase. Conclusions: Spanlastic gel might represent a promising carrier for the dermal delivery of TQ and AA for the management of hyperpigmentation conditions. Full article
(This article belongs to the Section Nanomedicine and Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Preparation of TQ and AA Spanlastics using the ethanol injection method.</p>
Full article ">Figure 2
<p>The developed spanlastics formula Z1 and Q4 containing AA and TQ, respectively.</p>
Full article ">Figure 3
<p>TEM images of spanlastics (<b>A</b>) Q-4 and (<b>B</b>) Z-1.</p>
Full article ">Figure 4
<p>FTIR Spectra of (<b>A</b>) Blank spanlastics, TQ, Q-4, and Physical mix and FTIR Spectra of (<b>B</b>) Blank spanlastics, AA, Z-1, and Physical mix.</p>
Full article ">Figure 5
<p>The prepared spanlastics R-1 gel and control gel (C-1).</p>
Full article ">Figure 6
<p>The flow curves (<b>A</b>) R-1 gel and C-1 gels determined at 32 °C. (<b>B</b>) The amplitude sweeps for R-1 gel. (<b>C</b>) The amplitude sweeps for C-1 gel. (<b>D</b>) The frequency sweeps for R-1 and C-1 gels. Data are presented as mean ± SD (n = 3).</p>
Full article ">Figure 7
<p>The release profile of Z-1, Q-4, TQ, and AA in R-1 gel. Data are presented as mean ± SD (n = 3).</p>
Full article ">Figure 8
<p>The Effect of R-1 gel and R-1 blank gel on HDF cell line viability compared to control, where no drug is added. Data are presented as mean ± SD (n = 3).</p>
Full article ">Figure 9
<p>The Tyrosinase Activity of TQ Spanlastics (Q-4), AA Spanlastics (Z-1), and their Combination Compared to Control, where no inhibitor is added. Data are presented as mean ± SD (n = 3).</p>
Full article ">
14 pages, 2614 KiB  
Article
Eco-Friendly Hydrogels Loading Polyphenols-Composed Biomimetic Micelles for Topical Administration of Resveratrol and Rutin
by Beatriz N. Guedes, Tatiana Andreani, M. Beatriz P. P. Oliveira, Faezeh Fathi and Eliana B. Souto
Biomimetics 2025, 10(1), 8; https://doi.org/10.3390/biomimetics10010008 - 27 Dec 2024
Viewed by 418
Abstract
In this study, we describe the development of hydrogel formulations composed of micelles loading two natural antioxidants—resveratrol and rutin—and the evaluation of the effect of a by-product on the rheological and textural properties of the developed semi-solids. This approach aims to associate the [...] Read more.
In this study, we describe the development of hydrogel formulations composed of micelles loading two natural antioxidants—resveratrol and rutin—and the evaluation of the effect of a by-product on the rheological and textural properties of the developed semi-solids. This approach aims to associate the advantages of hydrogels for topical administration of drugs and of lipid micelles that mimic skin composition for the delivery of poorly water-soluble compounds in combination therapy. Biomimetic micelles composed of L-α-phosphatidylcholine loaded with two distinct polyphenols (one non-flavonoid and one flavonoid) were produced using hot shear homogenisation followed by the ultrasonication method. All developed micelles were dispersed in a carbomer 940-based hydrogel to obtain three distinct semi-solid formulations, which were then characterised by analysing the thermal, rheological and textural properties. Olive pomace-based hydrogels were also produced to contain the same micelles as an alternative to respond to the needs of zero waste and circular economy. The thermograms showed no changes in the typical profiles of micelles when loaded into the hydrogels. The rheological analysis confirmed that the produced hydrogels achieved the ideal properties of a semi-solid product for topical administration. The viscosity values of the hydrogels loaded with olive pomace (hydrogels A) proved to be lower than the hydrogels without olive pomace (hydrogels B), with this ingredient having a considerable effect in reducing the viscosity of the final formulation, yet without compromising the firmness and cohesiveness of the gels. The texture analysis of both hydrogels A and B also exhibited the typical behaviour expected of a semi-solid system. Full article
(This article belongs to the Special Issue Advances in Biomaterials, Biocomposites and Biopolymers 2024)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of Resveratrol (reproduced after Chedea, Veronica Sanda et al. (2021) [<a href="#B27-biomimetics-10-00008" class="html-bibr">27</a>], under the terms and conditions of the Creative Commons Attribution (CC BY) license.</p>
Full article ">Figure 2
<p>Chemical structure of rutin (reproduced after Enogieru, Adaze Bijou et al. (2018) [<a href="#B28-biomimetics-10-00008" class="html-bibr">28</a>], under the terms and conditions of the Creative Commons Attribution (CC BY) license.</p>
Full article ">Figure 3
<p>Schematic representation of the production of resveratrol- and rutin-loaded micelles.</p>
Full article ">Figure 4
<p>Schematic representation of the production of resveratrol- and rutin-loaded micelles composed hydrogels.</p>
Full article ">Figure 5
<p>Hydrogels A (front and top) and hydrogels B (front and top).</p>
Full article ">Figure 6
<p>DSC analysis of the developed hydrogels and olive pomace.</p>
Full article ">Figure 7
<p>Oscillation frequency sweep test of Hydrogels A.</p>
Full article ">Figure 8
<p>Oscillation frequency sweep test of Hydrogels B.</p>
Full article ">
17 pages, 13237 KiB  
Article
Mechanical Decrosslinking and Reprocessing of Crosslinked Rotomolded Polypropylene Using Cryogenic-Assisted Shear Pulverization and Compression Molding
by Hibal Ahmad and Denis Rodrigue
Recycling 2024, 9(6), 129; https://doi.org/10.3390/recycling9060129 - 23 Dec 2024
Viewed by 408
Abstract
This paper presents a novel recycling approach for porous/foamed crosslinked rotomolded polypropylene (xPP) parts, originally designed for lightweight and thermal insulation. The method uses a cryogenic-assisted shear pulverization technique to produce parts by compression molding. The part’s final gel content and crosslink density [...] Read more.
This paper presents a novel recycling approach for porous/foamed crosslinked rotomolded polypropylene (xPP) parts, originally designed for lightweight and thermal insulation. The method uses a cryogenic-assisted shear pulverization technique to produce parts by compression molding. The part’s final gel content and crosslink density were found to depend on their dicumyl peroxide (DCP) content (0–2.5 phr) and characterized in terms of their chemical, thermal, physical and mechanical properties. The results show that this recycling technique allows for an effective reprocessing of the crosslinked materials since partial decrosslinking occurs. For example, the crosslink density decreased by 64% (3.10 to 1.11 × 10−3 mol/cm3) and the gel content by 9% (84.4% to 71.2%) at 2.5 phr DCP. Reprocessing through compression molding led to a compact and partially crosslinked structure resulting in significant improvements in terms of tensile strength (1480%), tensile modulus (604%), elongation at break (8900%), Shore A hardness (19%) and Shore D hardness (32%) compared to xPP samples (at 2.5 phr). This study paves the way for the development of more sustainable recycling methods, especially for crosslinked polymers, by providing new opportunities to reuse the wastes/end-of-life materials in advanced materials and different applications. Full article
Show Figures

Figure 1

Figure 1
<p>Possible chain mechanisms: (<b>a</b>) linear structure, (<b>b</b>) crosslinked structure, (<b>c</b>) disrupted structure after cryogenic shear pulverization and (<b>d</b>) reformed structure after compression molding (x = crosslinked and a,d = associative/dissociative links). Parts in blue are backbone and crosslinks are in red. The arrows refer to position with interaction/reaction.</p>
Full article ">Figure 2
<p>Structural characterization of the crosslinks in xPP [<a href="#B25-recycling-09-00129" class="html-bibr">25</a>] and r-xPP samples: (<b>a</b>) gel content and (<b>b</b>) crosslink density.</p>
Full article ">Figure 3
<p>FTIR spectra for typical xPP and rxPP samples at 1.0 phr DCP content.</p>
Full article ">Figure 4
<p>Density of the xPP and r-xPP [<a href="#B25-recycling-09-00129" class="html-bibr">25</a>] samples.</p>
Full article ">Figure 5
<p>Temperature-dependent thermal conductivity of the x-PP [<a href="#B25-recycling-09-00129" class="html-bibr">25</a>] and r-xPP samples at: (<b>a</b>) 23, (<b>b</b>) 40, (<b>c</b>) 60, (<b>d</b>) 80, (<b>e</b>) 100 and (<b>f</b>) 120 °C.</p>
Full article ">Figure 6
<p>Mechanical properties of the r-xPP samples: (<b>a</b>) tensile strength, (<b>b</b>) tensile modulus, (<b>c</b>) elongation at break and (<b>d</b>) hardness (Shore A and D).</p>
Full article ">Figure 7
<p>Schematic representation of the step-by-step process for recycling xPP parts using cryogenic pulverization and compression molding.</p>
Full article ">Figure 8
<p>Particles obtained: (<b>a</b>) without cryogenic treatment, (<b>b</b>) with cryogenic treatment, (<b>c</b>) sieved (1 mm) and (<b>d</b>) surface topology of the sieved particles (optical microscopy at 50×).</p>
Full article ">Figure 9
<p>Typical images of the recycled PP and xPP parts: (<b>a</b>) rPP, (<b>b</b>) 0.5, (<b>c</b>) 1.0, (<b>d</b>) 1.5, (<b>e</b>) 2.0 and (<b>f</b>) 2.5 phr r-xPP. The numbers represent the DCP content in the original samples.</p>
Full article ">
16 pages, 6552 KiB  
Article
An Ultra-Stable Polysaccharide Gel Plugging Agent for Water Shutoff in Mature Oil Reservoirs
by Yang Yang, Shuangxiang Ye, Ping Liu and Youqi Wang
Appl. Sci. 2024, 14(24), 11957; https://doi.org/10.3390/app142411957 - 20 Dec 2024
Viewed by 317
Abstract
Polyacrylamide-based gel plugging agents are extensively utilized in oilfields for water shutoff. However, their thermal stability, salt tolerance, and shear resistance are limited, making it difficult to achieve high-strength plugging and maintain stability under high-temperature and high-salinity reservoir conditions. This study proposes the [...] Read more.
Polyacrylamide-based gel plugging agents are extensively utilized in oilfields for water shutoff. However, their thermal stability, salt tolerance, and shear resistance are limited, making it difficult to achieve high-strength plugging and maintain stability under high-temperature and high-salinity reservoir conditions. This study proposes the use of chitosan (CTSs), a polysaccharide with a rigid cyclic structure, as the polymer. The organic cross-linker N,N′-methylenebisacrylamide (MBA) is incorporated via the Michael addition reaction mechanism to develop an ultra-stable, organically cross-linked chitosan gel system. The CTS/MBA gel system was evaluated under various environmental conditions using rheological testing and thermal aging to assess gel strength and stability. The results demonstrate significant improvements in gel strength and stability at high temperatures (up to 120 °C) and under high-shear conditions, as the increased cross-linking density enhanced resistance to thermal and mechanical degradation. Rapid gelation was observed with increasing MBA concentration, while pH and salinity further modulated gel properties. Scanning electron microscopy revealed the formation of a three-dimensional microstructure after gelation, which contributed to the enhanced properties. This study provides novel insights into optimizing polymer gel performance for the petroleum industry, particularly in high-temperature and high-shear environments. Full article
(This article belongs to the Special Issue Recent Advances and Emerging Technologies in Oil and Gas Production)
Show Figures

Figure 1

Figure 1
<p>Schematic of the gelation time determination.</p>
Full article ">Figure 2
<p>Effect of CTS concentration on gelation behavior (salinity: 30,000 mg/L; pH: 4.2; 120 °C).</p>
Full article ">Figure 3
<p>Effect of MBA concentration on gelation behavior (salinity: 30,000 mg/L; pH: 4.2; 120 °C).</p>
Full article ">Figure 4
<p>Effect of pH on gelation behavior (salinity: 30,000 mg/L; 120 °C).</p>
Full article ">Figure 5
<p>Effect of salinity on gelation behavior.</p>
Full article ">Figure 6
<p>Effect of temperature on the gel strength of the CTS/MBA gel system.</p>
Full article ">Figure 7
<p>Effect of temperature on the gelation time. (<b>a</b>) Salinity: 30,000 mg/L, pH: 4.2; (<b>b</b>) salinity: 30,000 mg/L, pH: 5.2; (<b>c</b>) salinity: 50,000 mg/L, pH: 4.2.</p>
Full article ">Figure 8
<p>Degradation rate at different shearing times (shear rate: 6000 r/min).</p>
Full article ">Figure 9
<p>Effect of shear degradation on the gelation time.</p>
Full article ">Figure 10
<p>Long-term aging gel strength variation (120 °C; salinity: 30,000 mg/L; pH: 4.2). (<b>a</b>) CTS/MBA gel base fluid; (<b>b</b>) aged for 30 days; (<b>c</b>) aged for 90 days.</p>
Full article ">Figure 11
<p>Microstructure of the CTS/MBA gel system (2.0 wt% CTS, 0.6 wt% MBA). (<b>a</b>) Before gelation; (<b>b</b>) after gelation at 120 °C.</p>
Full article ">Figure 12
<p>FTIR spectrum of all initial reactants and CTS/MBA.</p>
Full article ">Figure 13
<p>TGA curves of CTS (<b>a</b>) and CTS/MBA (<b>b</b>).</p>
Full article ">Figure 14
<p>(<b>a</b>) Chemical reaction during the gelling process. (<b>b</b>) Gelation mechanism of CTS/MBA.</p>
Full article ">
15 pages, 3747 KiB  
Article
Alginate Heterografted Copolymer Thermo-Induced Hydrogel Reinforced by PAA-g-P(boc-L-Lysine): Effects on Hydrogel Thermoresponsiveness
by Aikaterini-Ariadni Moschidi and Constantinos Tsitsilianis
Polymers 2024, 16(24), 3555; https://doi.org/10.3390/polym16243555 - 20 Dec 2024
Viewed by 533
Abstract
In this article, we report on the alginate heterografted by Poly(N-isopropyl acrylamide-co-N-tert-butyl acrylamide) and Poly(N-isopropyl acrylamide) (ALG-g-P(NIPAM86-co-NtBAM14)-g-PNIPAM) copolymer thermoresponsive hydrogel, reinforced by substituting part of the 5 wt% aqueous formulation by small amounts of Poly(acrylic acid)-g-P(boc-L-Lysine) (PAA-g-P(b-LL)) graft copolymer (up to 1 wt%). [...] Read more.
In this article, we report on the alginate heterografted by Poly(N-isopropyl acrylamide-co-N-tert-butyl acrylamide) and Poly(N-isopropyl acrylamide) (ALG-g-P(NIPAM86-co-NtBAM14)-g-PNIPAM) copolymer thermoresponsive hydrogel, reinforced by substituting part of the 5 wt% aqueous formulation by small amounts of Poly(acrylic acid)-g-P(boc-L-Lysine) (PAA-g-P(b-LL)) graft copolymer (up to 1 wt%). The resulting complex hydrogels were explored by oscillatory and steady-state shear rheology. The thermoresponsive profile of the formulations were affected remarkably by increasing the PAA-g-P(b-LL) component of the polymer blend. Especially, the sol-gel behavior altered to soft gel–strong gel behavior due to the formation of a semi-interpenetrating network based on the hydrophobic self-organization of the PAA-g-P(b-LL). In addition, the critical characteristics, namely Tc,thermothickening (temperature above which the viscosity increases steeply) and ΔT (transition temperature window), shifted and broadened to lower temperatures, respectively, due to the influence of the hydrophobic side chains P(b-LL) on the LCST of the PNIPAM-based grafted chains of the alginate. The effect of ionic strength was also examined, showing that this is another important factor affecting the thermoresponsiveness of the hydrogel. Again, the thermoresponsive profile of the hydrogel was changed significantly by the presence of salt. All the formulations showed self-healing capability and tolerance injectability, suitable for potential bioapplications in living bodies. Full article
(This article belongs to the Special Issue Advanced Study on Polymer-Based Hydrogels)
Show Figures

Figure 1

Figure 1
<p>Storage (G′) (closed) and loss (G″) (open) modulus (γ = 0.1%, 1 Hz) as a function of the temperature for the ALG-g-HG/PAA-P(b-LL) systems designated in wt% of the components, (<b>a</b>) 5/0 wt%, (<b>b</b>) 4.5/0.5 wt%, (<b>c</b>) 4.25/0.75 wt% and (<b>d</b>) 4/1 wt%. In the insets, the photos show the formulations at T = 15 °C (left down) and T = 50 °C (right up).</p>
Full article ">Figure 2
<p>T<sub>c,thermothickening</sub> and transition zone ΔΤ as a function of PAA-g-P(b-LL) component polymer concentration. The lines guide the eye.</p>
Full article ">Figure 3
<p>(<b>a</b>) Storage modulus (G′), and (<b>b</b>) tan(δ) as a function of temperature for the ALG-g-HG/PAA-g-P(b-LL) system at various compositions: 5/0 wt% (blue symbols); 4.5/0.5 wt% (pink symbols); 4.25/0.75 wt% (red symbols); 4/1 wt% PAA-g-P(b-LL) (green symbols). (<b>c</b>) Storage modulus (G′) at 20 and 50 °C, obtained from <a href="#polymers-16-03555-f003" class="html-fig">Figure 3</a>a as a function for the PAA-g-P(b-LL) component polymer concentration.</p>
Full article ">Figure 4
<p>G′ (<b>a</b>), η* (<b>b</b>) and critical strain, γ<sub>c</sub> (<b>d</b>) obtained from strain sweep data at 37 °C in the linear viscoelastic regime versus PAA-g-P(b-LL) percentage. (<b>c</b>) Strain sweep data of the ALG-g-HG/ PAA-g-P(b-LL) 4/1 wt% formulation.</p>
Full article ">Figure 5
<p>Time dependence of G′ (solid symbols) and G″ (open symbols) (1 Hz), subjected to consecutive variations in strain amplitude (as indicated), for the ALG-g-HG/PAA-g-P(b-LL) 4/1 wt% formulation at T = 37 °C and pH 7.4.</p>
Full article ">Figure 6
<p>Time dependence of G′ (solid symbols) and G″ (open symbols) (γ = 0.1%, 1 Hz) for (<b>a</b>,<b>c</b>) stepwise and (<b>b</b>,<b>d</b>) fast gradual temperature variation from 25 °C to 37 °C for the 4.5 wt% ALG-g-HG/0.5 wt% PAA-g-P(b-LL) (<b>a</b>,<b>b</b>) and 4 wt% ALG-g-HG/1 wt% PAA-g-P(b-LL) (<b>c</b>,<b>d</b>) formulations.</p>
Full article ">Figure 7
<p>Time dependence of apparent shear viscosity after consecutive stepwise variations of shear rates of the composite hydrogels at 18 °C.</p>
Full article ">Figure 8
<p>Storage modulus, G′ (<b>a</b>) and tan(δ) (<b>b</b>) (γ = 0.1%, 1 Hz) as a function of temperature for the 4.25 wt% ALG-g-HG/0.75 wt% PAA-g-P(b-LL) system at various salt NaCl concentrations: 0 M NaCl (blue symbols); 0.15M NaCl (red symbols); 0.3 M NaCl (pink symbols); 0.45 M NaCl (yellow symbols) (<b>c</b>) Salt concentration dependence of storage modulus (G′) at 18 and 50 °C, as obtained from Figure (<b>a</b>).</p>
Full article ">Figure 9
<p>(<b>a</b>) Apparent shear viscosity versus shear rate and (<b>b</b>) time dependence of apparent shear viscosity subjected stepwise to simultaneous variations in shear rate and temperature for the 4.25 wt% ALG-g-HG/0.75 wt% PAA-g-P(b-LL) formulation in the presence of 0.15 M NaCl. The red line (extrapolated by dotted line) in (<b>a</b>) is the linear fitting according to the double log transformation of Equation (2).</p>
Full article ">Scheme 1
<p>Schematic representation and chemical structures of the involved graft copolymer gelators.</p>
Full article ">
14 pages, 1700 KiB  
Article
Preparation of Green Tea Polyphenol-Loaded Diacylglycerol Nanostructured Lipid Carrier Hydrogels and Their Activities Related to Skin Protection
by Zhini Zhu, Qiu Xia, Xinxia Zhan, Wenyuan Li, Xuan He, Bo Wang, Qizhi Zhou, Jian Huang and Yong Ye
Materials 2024, 17(24), 6227; https://doi.org/10.3390/ma17246227 - 20 Dec 2024
Viewed by 423
Abstract
Diacylglycerol (DAG) is a functional oil but is rarely used in the cosmetic industry because low solubility, susceptibility to leakage and low viscosity to skin are still the main hurdles. A novel diacylglycerol nanostructured lipid carrier hydrogel (GTP-DAG-NLC-GEL) loaded with green tea polyphenol [...] Read more.
Diacylglycerol (DAG) is a functional oil but is rarely used in the cosmetic industry because low solubility, susceptibility to leakage and low viscosity to skin are still the main hurdles. A novel diacylglycerol nanostructured lipid carrier hydrogel (GTP-DAG-NLC-GEL) loaded with green tea polyphenol (GTP) was designed and successfully prepared to broaden DAG’s application in cosmetics, which significantly improved GTP stability and skin stickiness of DAG. The results showed that DAG-NLC-GEL had good viscosity, which was 980 Pa·s when the shear rate was 5 rpm, and its viscosity decreased quickly with the increase in shear rate, making it easily expand on skin. Meanwhile, the encapsulation rate and drug loading of GTP in GDP-DAG-NLC-GEL reached 86.7% and 2.6%, respectively, and the DPPH free radicals scavenging rate and inhibition rate of the advanced glycation end-products (AGEs) were 85.46% and 89.72%, respectively, which indicate that GTP-DAG-NLC-GEL has significant skin sunscreen, antioxidant and anti-glycation activities. The GTP-loaded nanostructured lipid carrier hydrogel can be deemed to have great prospects for skin protection in cosmetics. Full article
Show Figures

Figure 1

Figure 1
<p>TEM images of (<b>a</b>) DAG-NLC, (<b>b</b>) GTP-DAG-NLC, (<b>c</b>) GTP-DAG-NLC-GEL.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern of GTP-DAG-NLC-GEL and GTP. (<b>b</b>) FT-IR spectra of GTP-DAG-NLC-GEL and GTP. (<b>c</b>) Thermogravimetric curves of GTP-DAG-NLC-GEL and GTP. (<b>d</b>) DSC curves of octanoic acid diglyceride and its DAG-NLC. (<b>e</b>) DSC curves of GTP-DAG-NLC and GTP-DAG-NLC-GEL.</p>
Full article ">Figure 2 Cont.
<p>(<b>a</b>) XRD pattern of GTP-DAG-NLC-GEL and GTP. (<b>b</b>) FT-IR spectra of GTP-DAG-NLC-GEL and GTP. (<b>c</b>) Thermogravimetric curves of GTP-DAG-NLC-GEL and GTP. (<b>d</b>) DSC curves of octanoic acid diglyceride and its DAG-NLC. (<b>e</b>) DSC curves of GTP-DAG-NLC and GTP-DAG-NLC-GEL.</p>
Full article ">Figure 3
<p>Viscosity of GTP-DAG-NLC-GEL at different shear rates.</p>
Full article ">Figure 4
<p>(<b>a</b>) Standard curve of absorbance vs. concentration of GTP. (<b>b</b>) The cumulative release curves of GTP-DAG-NLC and GTP-DAG-NLC-GEL.</p>
Full article ">Figure 5
<p>Skin protection activity of GTP-DAG-NLC-GEL: (<b>a</b>) UV-vis absorbance spectra. (<b>b</b>) DPPH free radical scavenging rate. (<b>c</b>) Inhibition rate of AGE.</p>
Full article ">Figure 5 Cont.
<p>Skin protection activity of GTP-DAG-NLC-GEL: (<b>a</b>) UV-vis absorbance spectra. (<b>b</b>) DPPH free radical scavenging rate. (<b>c</b>) Inhibition rate of AGE.</p>
Full article ">
22 pages, 10538 KiB  
Article
Changes in Functional Properties and In Vitro Digestibility of Black Tartary Buckwheat Starch by Autoclaving Combination with Pullulanase Treatment
by Faying Zheng, Fuxin Nie, Ye Qiu, Yage Xing, Qinglian Xu, Jianxiong Chen, Ping Zhang and Hong Liu
Foods 2024, 13(24), 4114; https://doi.org/10.3390/foods13244114 - 19 Dec 2024
Viewed by 598
Abstract
The processing properties of resistant starch (RS) and its digestion remain unclear, despite the widespread use of autoclaving combined with debranching in its preparation. In this study, the physicochemical, rheological and digestibility properties of autoclaving modified starch (ACB), autoclaving–pullulanase modified starch (ACPB) and [...] Read more.
The processing properties of resistant starch (RS) and its digestion remain unclear, despite the widespread use of autoclaving combined with debranching in its preparation. In this study, the physicochemical, rheological and digestibility properties of autoclaving modified starch (ACB), autoclaving–pullulanase modified starch (ACPB) and native black Tartary buckwheat starch (NB) were compared and investigated. The molecular weight and polydispersity index of modified starch was in the range of 0.15 × 104~1.90 × 104 KDa and 1.88~2.82, respectively. In addition, the SEM results showed that both modifications influenced the morphological characteristics of the NB particles, and their particles tended to be larger in size. Autoclaving and its combination with pullulanase significantly increased the short-range ordered degree, resistant starch yield and water- and oil-absorption capacities, and decreased the syneresis properties with repeated freezing/thawing cycles. Moreover, rheological analysis showed that both ACB and ACPB exhibited shear-thinning behavior and lower gel elasticity as revealed by the power law model and steady-state scan. The degradation of starch chains weakened the interaction of starch molecular chains and thus changed the gel network structure. The in vitro digestion experiments demonstrated that ACB and ACPB exhibited greater resistance to enzymatic digestion compared to the control, NB. Notably, the addition of pullulanase inhibited the hydrolysis of the ACB samples, and ACPB showed greater resistance against enzymatic hydrolysis. This study reveals the effects of autoclaving combined with debranching on the processing properties and functional characteristics of black Tartary buckwheat starch. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM micrographs of NB (<b>A1</b>,<b>A2</b>,<b>A3</b>), ACB (<b>B1</b>,<b>B2</b>,<b>B3</b>) and ACPB (<b>C1</b>,<b>C2</b>,<b>C3</b>) (magnification: 500×, 1000× and 2000×). NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment.</p>
Full article ">Figure 2
<p>Particle size distribution of NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment.</p>
Full article ">Figure 3
<p>FTIR spectra of NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment.</p>
Full article ">Figure 4
<p>Water solubility (<b>A</b>), swelling power (<b>B</b>), water absorbing capacity (<b>C</b>) and oil absorbing capacity (<b>D</b>) of NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment. For (<b>A</b>,<b>B</b>), the means of different letters at the same temperature are significantly different at <span class="html-italic">p</span> &lt; 0.05. For (<b>C</b>,<b>D</b>), the means with different letters are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Light transmittance of NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment. Data with different letters at the same test time are significantly different at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Steady shear flow data. Steady-state shear viscosity (η) (<b>A</b>) and shear stress (τ) (<b>B</b>) for NB, ACB and ACPB. γ: shear rate; NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment.</p>
Full article ">Figure 7
<p>Dynamic frequency sweep data. Storage modulus (G′), loss modulus (G″) (<b>A</b>) and loss tangent (tanδ) (<b>B</b>) for NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment.</p>
Full article ">Figure 8
<p>Digestibility of NB, ACB and ACPB. Hydrolysis curve (<b>A</b>); rapidly digestible starch (RDS), slowly digestible starch (SDS), and resistant starch (RS) of samples (<b>B</b>). NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment. ns: the means are not significantly different at <span class="html-italic">p</span> &gt; 0.05 with each other. *: the means are significantly different at <span class="html-italic">p</span> &lt; 0.05 with each other.</p>
Full article ">Figure 9
<p>Fit data of first-order kinetics of NB, ACB and ACPB. NB, black Tartary buckwheat native starch; ACB, black Tartary buckwheat starch subjected to autoclaving treatment; ACPB, black Tartary buckwheat starch subjected to autoclaving–pullulanase combined treatment. C (%) is the total amount of starch digested at digestion time t; C∞(%) is the total amount of starch digested at 240 min.</p>
Full article ">
21 pages, 5088 KiB  
Article
Formation and Characterization of Mycelium–Potato Protein Hybrid Materials for Application in Meat Analogs or Substitutes
by Ramdattu Santhapur, Disha Jayakumar and David Julian McClements
Foods 2024, 13(24), 4109; https://doi.org/10.3390/foods13244109 - 19 Dec 2024
Viewed by 1114
Abstract
There is increasing interest in the development of meat analogs due to growing concerns about the environmental, ethical, and health impacts of livestock production and consumption. Among non-meat protein sources, mycoproteins derived from fungal fermentation are emerging as promising meat alternatives because of [...] Read more.
There is increasing interest in the development of meat analogs due to growing concerns about the environmental, ethical, and health impacts of livestock production and consumption. Among non-meat protein sources, mycoproteins derived from fungal fermentation are emerging as promising meat alternatives because of their natural fibrous structure, high nutritional content, and low environmental impact. However, their poor gelling properties limit their application in creating meat analogs. This study investigated the potential of creating meat analogs by combining mycoprotein (MCP), a mycelium-based protein, with potato protein (PP), a plant-based protein, to create hybrid products with meat-like structures and textures. The PP-MCP composites were evaluated for their physicochemical, rheological, textural, and microstructural properties using electrophoresis, differential scanning calorimetry, dynamic shear rheology, texture profile analysis, confocal fluorescence microscopy, and scanning electron microscopy analyses. The PP-MCP hybrid gels were stronger and had more fibrous structures than simple PP gels, which was mainly attributed to the presence of hyphae fibers in mycelia. Dynamic shear rheology showed that the PP-MCP hybrids formed irreversible heat-set gels with a setting temperature of around 70 °C during heating, which was attributed to the unfolding and aggregation of the potato proteins. Confocal and electron microscopy analyses showed that the hybrid gels contained a network of mycelia fibers embedded within a potato protein matrix. The hardness of the PP-MCP composites could be increased by raising the potato protein content. These findings suggest that PP-MCP composites may be useful for the development of meat analogs with more meat-like structures and textures. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Impact of pH on the zeta potential of 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) MCP, 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) MCP + PP (1:1) and 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) PP dispersions.</p>
Full article ">Figure 2
<p>pH-dependence of the aggregation of the 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) of MCP, 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) PP and 0.1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) 1:1 MCP + PP samples.</p>
Full article ">Figure 3
<p>Differential scanning calorimetry profile of 15 wt% potato protein (PP) and 15 wt% mycoprotein (MCP) during heating from 25 to 100 °C with at 3 °C/min having denaturation temperature (T<sub>d</sub>) of 65.53 °C indicated with arrow.</p>
Full article ">Figure 4
<p>Temperature sweep results of (<b>a</b>) 15% potato protein (PP); (<b>b</b>) 15% mycoprotein (MCP); and (<b>c</b>) 15% MCP and 15% PP. The storage (G′) and loss (G″) moduli of the samples were measured as they were heated from 25 to 90 °C, held at 90 °C, and then cooled from 90 to 10 °C (strain = 0.1, frequency = 1 Hz). The red arrows show heating, while the blue arrows show cooling. For most temperatures, G′ &gt; G″ for all samples, indicating they were predominantly elastic-like materials.</p>
Full article ">Figure 5
<p>Frequency sweep results of 15% potato protein (PP), 15% mycoprotein (MCP), and 15% mycoprotein-15% potato protein (15% MCP + 15% PP) hybrid gels. The storage modulus (G′) and loss modulus (G″) of the samples were measured as the frequency was increased at a strain of 0.1% and 25 °C. For all frequencies, G′ &gt; G″ for all samples, indicating they were predominantly elastic-like materials.</p>
Full article ">Figure 6
<p>Strain sweep results of 15% pure potato protein,15% mycoprotein and15% potato protein-mycoprotein hybrid gels. The complex shear modulus (G*) of the samples were measured as the strain was increased from 0.01% to 1000% at 25 °C.</p>
Full article ">Figure 7
<p>Effect of single composition on the stress–strain relationship of Potato protein and potato protein-mycelium hybrids during single compression-decompression experiments (25 °C).</p>
Full article ">Figure 8
<p>Double compression curves of 15% potato protein (15% PP) and 15% mycoprotein + 15% potato protein (15% MCP + 15% PP) hybrid gels. The force versus time curves were measured at 25 °C with 50% final strain and 2 mm/s pre-test, test, and post-test speeds.</p>
Full article ">Figure 9
<p>Confocal microscopy images of 15% mycoprotein (15% MCP), 15% potato protein (15% PP) and 15% mycoprotein + 15% potato protein (15% MCP + 15% PP) hybrid gels (25 °C). The images of the pure mycoprotein samples show they contained fibrous structures (stained blue and green), which were presumably chitin- and protein-rich hyphae. The images of the pure potato proteins showed that they contained large protein aggregates (stained dark green) dispersed in a protein-rich aqueous phase (stained light green). The black regions were probably holes formed during sample preparation. The images of the hybrid samples showed that they contained some fibrous structures (stained blue and green), which were presumably chitin- and protein-rich hyphae, distributed in a protein-rich network (stained green).</p>
Full article ">Figure 10
<p>The scanning electron microscopy images of 15% mycoprotein, 15% potato protein and 15% mycoprotein + 15% potato protein hybrid gel. Scale bars are 100 μm for 300× and 10 μm for 3000× and 2000× for 15% potato protein. The digital photographs show the overall appearance of the samples before freeze drying.</p>
Full article ">
15 pages, 3187 KiB  
Article
Liquid Crystal/Carbon Nanotube/Polyaniline Composites and Their Coating Orientation Patterning Applications
by Fuqiang Chu, Haikuo Zhang, Xu Zhou, Yuhang Fu, Hang Dong, Shuo Wang, Jilei Chao and Xin Wang
Coatings 2024, 14(12), 1568; https://doi.org/10.3390/coatings14121568 - 13 Dec 2024
Viewed by 717
Abstract
In this work, a coating method was used to prepare a liquid crystal physical gel with a high orientation of liquid crystal molecules, excellent electrical conductivity, and mechanical stability. The liquid crystal matrix used was nematic phase liquid crystal (5CB), the gel factor [...] Read more.
In this work, a coating method was used to prepare a liquid crystal physical gel with a high orientation of liquid crystal molecules, excellent electrical conductivity, and mechanical stability. The liquid crystal matrix used was nematic phase liquid crystal (5CB), the gel factor was polyvinyl alcohol (PVA), and the conductive filler was carbon nanotubes/polyaniline (CNT/PANI). Chemical in situ polymerization was used to create CNT/PANI composites, wherein polyaniline encapsulates the carbon nanotubes to enhance their dispersion. At 4 mm/s, 7.2 N of coating pressure, and 72 s of interval duration, the shear flow-induced orientation was achieved. The consistent and large-area orientation of the liquid crystal molecules was realized and the orientation direction of the liquid crystal molecules was parallel to the coating direction. Additionally, a type of stress sensor assembly based on multiple coating demonstrated a good sensor performance in the 90° bending test and high sensitivity in the 20% tensile test, with a sensor sensitivity of 23.25. Regarding the use of liquid crystal materials in flexible electronic devices, it is quite important. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Preparation process of carbon nanotube/polyaniline composites. (<b>b</b>) Preparation process of liquid crystal/carbon nanotube/polyaniline physical gel (<b>c</b>) Three different physical gels (<b>d</b>) Coating orientation process of liquid crystals in liquid crystal/carbon nanotube/polyaniline physical gels. (<b>e</b>) Sensor Preparation and Assembly Process (<b>f</b>) Liquid crystal/carbon nanotube/polyaniline conductive film-assembled sensors.</p>
Full article ">Figure 2
<p>SEM images of the (<b>a</b>) CNT (<b>b</b>) CNT/PANI (<b>c</b>) FTIR images of the CNT, PANI, CNT/PANI (<b>d</b>) DSC images of the LC, PVA, LC/PVA, LC/CNT/PVA, LC/CNT/PANI/PVA.</p>
Full article ">Figure 3
<p>POM plots of LC microdroplets in LC/CNT physical gels after coating orientation in transmission mode. POM plots of LC micro droplets with different angles of the cross-polarizer: (<b>a</b>) +45°, (<b>b</b>) 0°, (<b>c</b>) −45°. (<b>d</b>) POM plot of an LC micro droplet at +45° to the cross-polarizer with the λ wavelength compensator inserted (550 nm delay and its slow axis at +45° to the cross-polarizer). (<b>e</b>) Light maps of LC microdroplets. (<b>f</b>) POM plot at −45° to the cross-polarizer with the λ-wavelength compensator inserted and the direction of the arrow indicating the coating.</p>
Full article ">Figure 4
<p>(<b>a</b>) Raman spectra of unoriented LC/PVA film, coated oriented LC/PVA film, coated oriented LC/CNT/PVA film, and coated oriented LC/CNT/PANI/PVA film. (<b>b</b>) Schematic diagram of coating orientation. (<b>c</b>) Effect of coating pressure on liquid crystal orientation (<b>d</b>) Effect of coating interval time on liquid crystal orientation (<b>e</b>) Effect of coating speed on liquid crystal orientation.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM images of the coated oriented LC/CNT/PANI conductive film (<b>b</b>) Electrochemical AC impedance mapping of four different conductive films.</p>
Full article ">Figure 6
<p>(<b>a</b>) Resistance change at different strains (<b>b</b>) Resistance change curve at 20% strain (<b>c</b>) Resistance change curve during stretching at 20% strain (<b>d</b>) Resistance change at different bending angle (<b>e</b>) Relative resistance change at 90° bending angle (<b>f</b>) Resistance change curve during stretching at 90° bending angle.</p>
Full article ">Figure 7
<p>Relative resistance of finger joint flexion.</p>
Full article ">
14 pages, 6222 KiB  
Article
Rheological Properties of Emulsions Stabilized by Cellulose Derivatives with the Addition of Ethyl Alcohol
by Sylwia Różańska, Jacek Różański, Patrycja Wagner and Ewelina Warmbier-Wytykowska
Materials 2024, 17(24), 6090; https://doi.org/10.3390/ma17246090 - 13 Dec 2024
Viewed by 509
Abstract
The paper presents the results of research on the rheological properties and stability of oil-in-water emulsions containing cellulose derivatives: methylcellulose, hydroxyethylcellulose, and hydroxypropylmethylcellulose. The continuous phase of the emulsion was a 70% ethanol (EtOH) solution by volume. The dispersed phase consisted of mineral, [...] Read more.
The paper presents the results of research on the rheological properties and stability of oil-in-water emulsions containing cellulose derivatives: methylcellulose, hydroxyethylcellulose, and hydroxypropylmethylcellulose. The continuous phase of the emulsion was a 70% ethanol (EtOH) solution by volume. The dispersed phase consisted of mineral, linseed, and canola oils (20% by volume). Rheological measurements were performed in both steady and oscillatory flow. Emulsion stability was assessed on visual observation and changes in droplet diameter over a period of 5 months after preparation. Relatively stable emulsions were obtained without the addition of low-molecular-weight surfactants, exhibiting viscoelastic properties. The presence of ethanol in the continuous phase significantly slowed down the processes of emulsion sedimentation or creaming, as well as droplet coalescence. The reasons for the slow phase separation were linked to changes in density and zero-shear viscosity of the continuous phase caused by the addition of EtOH. All emulsions were highly polydisperse, and the addition of methylcellulose and hydroxypropylmethylcellulose further led to the formation of strongly flocculated emulsions. Droplet flocculation resulted in highly viscoelastic fluids. In particular, for emulsions containing hydroxypropylmethylcellulose, the ratio of the storage modulus to the loss modulus approached a value close to 0.1, which is characteristic of gels. Full article
(This article belongs to the Section Polymeric Materials)
Show Figures

Figure 1

Figure 1
<p>Sample photos of the emulsion taken 45 days after preparation: (<b>a</b>) HEC 1.5% wt./EtOH, (<b>b</b>) MC 1.5% wt./EtOH, (<b>c</b>) HPMC 0.7% wt./EtOH.</p>
Full article ">Figure 2
<p>Viscosity curves for aqueous polymer solutions and polymers in water/ethanol mixtures (70% vol. EtOH).</p>
Full article ">Figure 3
<p>G′ and G″ moduli (<b>a</b>) and tan δ = G″/G′ (<b>b</b>) as a function of oscillation frequency ω for polymers in water/ethanol mixtures.</p>
Full article ">Figure 4
<p>G′ and G″ moduli and tan δ = G″/G′ as a function of oscillation frequency ω for emulsions with (<b>a</b>) linseed oil, (<b>b</b>) canola oil, (<b>c</b>) mineral oil.</p>
Full article ">Figure 5
<p>Sample photos of emulsions with added polymers/EtOH mixtures and various oils after preparation: (<b>a</b>) HEC/EtOH—canola oil, HEC 1.5% wt., (<b>b</b>) HEC/EtOH—linseed oil, HEC 1.5% wt., (<b>c</b>) HEC/EtOH—mineral oil, HEC 1.5% wt., (<b>d</b>) MC/EtOH—canola oil, MC 1.5% wt., (<b>e</b>) MC/EtOH—linseed oil, MC 1.5% wt., (<b>f</b>) MC/EtOH—mineral oil, MC 1.5% wt., (<b>g</b>) HPMC/EtOH—canola oil, HPMC 0.7% wt., (<b>h</b>) HPMC/EtOH—linseed oil, HPMC 0.7% wt., (<b>i</b>) HPMC/EtOH—mineral oil, HPMC 0.7% wt.</p>
Full article ">Figure 6
<p>Comparison of the dependence of complex viscosity on oscillation frequency for emulsions containing linseed oil (LO), canola oil (CO), and mineral oil (MO) (emulsions with the addition of HEC (2% wt.) and EtOH (70% vol.)).</p>
Full article ">
16 pages, 2939 KiB  
Article
Extraction Methods and Characterization of β-Glucans from Yeast Lees of Wines Produced Using Different Technologies
by Ana Chioru, Aurica Chirsanova, Adriana Dabija, Ionuț Avrămia, Alina Boiştean and Ancuța Chetrariu
Foods 2024, 13(24), 3982; https://doi.org/10.3390/foods13243982 - 10 Dec 2024
Viewed by 1056
Abstract
Wine lees, the second most significant by-product of winemaking after grape pomace, have received relatively little attention regarding their potential for valorization. Despite their rich content in bioactive components such as β-glucans, industrial utilization faces challenges, particularly due to variability in their composition. [...] Read more.
Wine lees, the second most significant by-product of winemaking after grape pomace, have received relatively little attention regarding their potential for valorization. Despite their rich content in bioactive components such as β-glucans, industrial utilization faces challenges, particularly due to variability in their composition. This inconsistency impacts the reliability and standardization of final products, limiting broader adoption in industrial applications. β-Glucans are dietary fibers or polysaccharides renowned for their diverse bioactive properties, including immunomodulatory, antioxidant, anti-inflammatory, antitumor, and cholesterol- and glucose-lowering effects. They modulate the immune system by activating Dectin-1 and TLR receptors on immune cells, enhancing phagocytosis, cytokine production, and adaptive immune responses. Their antioxidant activity arises from neutralizing free radicals and reducing oxidative stress, thereby protecting cells and tissues. β-Glucans also exhibit antitumor effects by inhibiting cancer cell growth, inducing apoptosis, and preventing angiogenesis, the formation of new blood vessels essential for tumor development. Additionally, they lower cholesterol and glucose levels by forming a viscous gel in the intestine, which reduces lipid and carbohydrate absorption, improving metabolic health. The biological activity of β-glucans varies with their molecular weight and source, further highlighting their versatility and functional potential. This study investigates how grape variety, vinification technology and extraction methods affect the yield and properties of β-glucans extracted from wine lees. The physico-chemical and mineral composition of different wine lees were analyzed, and two extraction methods of β-glucans from wine lees were tested: acid-base extraction and autolysis. These two methods were also tested under ultrasound-assisted conditions at different frequencies, as well as without the use of ultrasound. The β-glucan yield and properties were evaluated under different conditions. FTIR spectroscopy was used to assess the functional groups and structural characteristics of the β-glucans extracted from the wine lees, helping to confirm their composition and quality. Rheological behavior of the extracted β-glucans was also assessed to understand the impact of extraction method and raw material origin. The findings highlight that vinification technology significantly affects the composition of wine lees, while both the extraction method and yeast origin influence the yield and type of β-glucans obtained. The autolysis method provided higher β-glucan yields (18.95 ± 0.49% to 39.36 ± 0.19%) compared to the acid–base method (3.47 ± 0.66% to 19.76 ± 0.58%). FTIR spectroscopy revealed that the β-glucan extracts contain a variety of glucan and polysaccharide types, with distinct β-glucans (β-1,4, β-1,3, and β-1,6) identified through specific absorption peaks. The rheological behavior of suspensions exhibited pseudoplastic or shear-thinning behavior, where viscosity decreased significantly as shear rate increased. This behavior, observed across all β-glucan extracts, is typical of polymer-containing suspensions. These insights are critical for optimizing β-glucan extraction processes, supporting sustainability efforts and waste valorization in the wine industry. Efficient extraction of β-glucans from natural sources like wine lees offers a promising path toward their industrial application as valuable functional compounds. Full article
Show Figures

Figure 1

Figure 1
<p>Microscopic images of winery yeast lees (100× magnification objective). (<b>a</b>) Semi-dry white wine (SVAM), (<b>b</b>) sweet white wine (SVR), (<b>c</b>) dry red wine (SVRS), (<b>d</b>) sparkling white wine (SVS).</p>
Full article ">Figure 2
<p>FT−IR−ATR spectra of the most representative samples.</p>
Full article ">Figure 3
<p>Shear rate dependence of viscosity for 2% suspension of β-glucans extracted by the acid–base method.</p>
Full article ">Figure 4
<p>Shear rate dependence of viscosity for 2% suspension of β-glucans extracted by autolysis.</p>
Full article ">
Back to TopTop