[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (370)

Search Parameters:
Keywords = GelMA

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 6796 KiB  
Article
Whole-Cell Vaccine Preparation Through Prussian Blue Nanoparticles-Elicited Immunogenic Cell Death and Loading in Gel Microneedles Patches
by Wenxin Fu, Qianqian Li, Jingyi Sheng, Haoan Wu, Ming Ma and Yu Zhang
Gels 2024, 10(12), 838; https://doi.org/10.3390/gels10120838 - 19 Dec 2024
Abstract
Tumor whole-cell vaccines are designed to introduce a wide range of tumor-associated antigens into the body to counteract the immunosuppression caused by tumors. In cases of lymphoma of which the specific antigen is not yet determined, the tumor whole-cell vaccine offers distinct advantages. [...] Read more.
Tumor whole-cell vaccines are designed to introduce a wide range of tumor-associated antigens into the body to counteract the immunosuppression caused by tumors. In cases of lymphoma of which the specific antigen is not yet determined, the tumor whole-cell vaccine offers distinct advantages. However, there is still a lack of research on an effective preparation method for the lymphoma whole-cell vaccine. To solve this challenge, we prepared a whole-cell vaccine derived from non-Hodgkin B-cell lymphoma (A20) via the photothermal effect mediated by Prussian blue nanoparticles (PBNPs). The immune activation effect of this vaccine against lymphoma was verified at the cellular level. The PBNPs-treated A20 cells underwent immunogenic cell death (ICD), causing the loss of their ability to form tumors while retaining their ability to trigger an immune response. A20 cells that experienced ICD were further ultrasonically crushed to prepare the A20 whole-cell vaccine with exposed antigens and enhanced immunogenicity. The A20 whole-cell vaccine was able to activate the dendritic cells (DCs) to present antigens to T cells and trigger specific immune responses against lymphoma. Whole-cell vaccines are primarily administered through direct injection, a method that often results in low delivery efficiency and poor patient compliance. Comparatively, the microneedle patch system provides intradermal delivery, offering enhanced lymphatic absorption and improved patient adherence due to its minimally invasive approach. Thus, we developed a porous microneedle patch system for whole-cell vaccine delivery using Gelatin Methacryloyl (GelMA) hydrogel and n-arm-poly(lactic-co-glycolic acid) (n-arm-PLGA). This whole-cell vaccine combined with porous gel microneedle patch delivery system has the potential to become a simple immunotherapy method with controllable production and represents a promising new direction for the treatment of lymphoma. Full article
(This article belongs to the Special Issue Gel-Based Drug Delivery Systems for Cancer Treatment (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Temperature changes and final temperature of A20 photothermal system with or without Prussian blue nanoparticles (PBNPs) under different laser power densities (LPDs). (<b>b</b>) Flow cytometric scatter diagrams of A20 cells after different treatments. (<b>c</b>) Changes in the optical microstructure of cells before and after photothermal treatment (808 nm laser, 0.75 W/cm<sup>2</sup>) in A20. (<b>d</b>) The survival and death rates of A20 cells following photothermal treatment with different LPDs.</p>
Full article ">Figure 2
<p>(<b>a</b>) ATP content secreted by A20 cells after photothermal treatment with different LPDs. (<b>b</b>) Flow cytometry characterization results of high mobility group protein B1 (HMGB1) content in A20 cytoplasm after photothermal treatment with different LPDs. (<b>c</b>) Flow cytometry characterization results of calreticulin (CRT) content in A20 cell membrane after photothermal treatment with different LPDs. (<b>d</b>) Confocal fluorescence images of HMGB1 content in A20 cytoplasm treated with different LPDs (20×). (<b>e</b>) Confocal fluorescence images of CRT content in A20 cell membrane treated with different LPDs (20×). *** <span class="html-italic">p</span>  &lt;  0.001, **** <span class="html-italic">p</span>  &lt;  0.0001, ns: not significant.</p>
Full article ">Figure 3
<p>(<b>a</b>) ATP content secreted by A20 cells after different heat treatments. (<b>b</b>) Flow cytometry characterization results of HMGB1 in A20 cytoplasm after different heat treatments. (<b>c</b>) Flow cytometry characterization results of CRT expression on A20 cell membrane after different heat treatments. (<b>d</b>) Confocal fluorescence images of HMGB1 content in A20 cytoplasm after different heat treatments (20×). (<b>e</b>) Confocal fluorescence images of CRT expression on A20 cell membrane after different heat treatments (20×). **** <span class="html-italic">p</span>  &lt;  0.0001.</p>
Full article ">Figure 4
<p>(<b>a</b>) Particle size distribution of A20 that underwent ICD (A20–ICD) ultrasonic fragmentation products under different ultrasound powers. (<b>b</b>) Protein content of A20–ICD and its ultrasonic fragmentation products. (<b>c</b>) transmission electron microscope (TEM) image (100×) of A20–ICD lysate products (red arrow indicates) within 1 h after ultrasonic fragmentation. (<b>d</b>) TEM image (100×) of A20–ICD lysate products (red arrow indicates) 10 days after ultrasonic fragmentation. ns: not significant.</p>
Full article ">Figure 5
<p>(<b>a</b>) Activation rate of bone marrow dendritic cells (BMDCs) stimulated with different treatments. (<b>b</b>) Activation rate of T cells stimulated with different treatments. (<b>c</b>) The proliferative effect of T cells stimulated with different treatments. *** <span class="html-italic">p</span>  &lt;  0.001, **** <span class="html-italic">p</span>  &lt;  0.0001, ns: not significant.</p>
Full article ">Figure 6
<p>Poly(lactic-co-glycolic acid) (PLGA) porous microneedle patch (<b>a</b>) Top view photo. (<b>b</b>) Side view photo. (<b>c</b>) Surface mounted microneedle array scanning electron microscope (SEM) image (100×). (<b>d</b>) SEM image of microneedle body (500×). (<b>e</b>) SEM image of surface pore structure of microneedle body (5 k×).</p>
Full article ">Figure 7
<p>Construction principle of PLGA porous microneedles with Gelatin Methacryloyl (GelMA) hydrogel.</p>
Full article ">Figure 8
<p>Loading of porous microneedle system for A20 whole-cell vaccine (10×, scale bar: 100 μm, scanning of confocal layer from 0 to 600 μm indicating microneedle bottom to tip).</p>
Full article ">Figure 9
<p>(<b>a</b>) Porous microneedle system release of A20 whole-cell vaccine. (<b>b</b>) First order equation release model fitting of porous microneedle patch system for whole-cell vaccine. **** <span class="html-italic">p</span>  &lt;  0.0001, ns: not significant.</p>
Full article ">
15 pages, 9729 KiB  
Article
Microstructure and Bioactivity of Ca- and Mg-Modified Silicon Oxycarbide-Based Amorphous Ceramics
by Qidong Liu, Hongmei Chen, Xiumei Wu, Junjie Yan, Biaobiao Yang, Chenying Shi, Yunping Li and Shu Yu
Materials 2024, 17(24), 6159; https://doi.org/10.3390/ma17246159 - 17 Dec 2024
Viewed by 226
Abstract
Silicon oxycarbide (SiOC), Ca- and Mg-modified silicon oxycarbide (SiCaOC and SiMgOC) were synthesized via sol–gel processing with subsequent pyrolysis in an inert gas atmosphere. The physicochemical structures of the materials were characterized by XRD, SEM, FTIR, and 29Si MAS NMR. Biocompatibility and [...] Read more.
Silicon oxycarbide (SiOC), Ca- and Mg-modified silicon oxycarbide (SiCaOC and SiMgOC) were synthesized via sol–gel processing with subsequent pyrolysis in an inert gas atmosphere. The physicochemical structures of the materials were characterized by XRD, SEM, FTIR, and 29Si MAS NMR. Biocompatibility and in vitro bioactivity were detected by MTT, cell adhesion assay, and simulated body fluid (SBF) immersion test. Mg and Ca were successfully doped into the network structure of SiOC, and the non-bridging oxygens (NBO) were formed. The hydroxycarbonate apatite (HCA) was formed on the modified SiOC surface after soaking in simulated body fluid (SBF) for 14 days, and the HCA generation rate of SiCaOC was higher than that of SiMgOC. Accompanying the increase of bioactivity, the network connectivity (NC) of the modified SiOC decreased from 6.05 of SiOC to 5.80 of SiCaOC and 5.60 of SiMgOC. However, structural characterization and biological experiments revealed the nonlinear relationship between the biological activity and NC of the modified SiOC materials. Full article
Show Figures

Figure 1

Figure 1
<p>XRD analysis of different samples after 1000 °C heat treatment.</p>
Full article ">Figure 2
<p>SEM and EDS analysis of different materials: (<b>a</b>) SiOC glass; (<b>b</b>) SiCaOC glass; (<b>c</b>) SiMgOC glass.</p>
Full article ">Figure 3
<p>FTIR spectra of SiOC, SiCaOC, and SiMgOC glasses.</p>
Full article ">Figure 4
<p><sup>29</sup>Si MAS NMR of three materials: (<b>a</b>) SiOC glass; (<b>b</b>) SiCaOC glass; (<b>c</b>) SiMgOC glass. The experimental (grey line) and simulated (red line) spectra, as well as the individual simulation components (black lines), are shown. The results of the simulation correspond to Q<sub>4</sub>SiO<sub>4</sub> (I), Q<sub>3</sub>SiO<sub>4</sub> (II), SiO<sub>3</sub>C (III), SiO<sub>2</sub>C<sub>2</sub> (IV), SiOC<sub>3</sub> (V), and SiC<sub>4</sub> (VI).</p>
Full article ">Figure 5
<p>The surface morphology of the samples immersed in SBF for different immersion times: (<b>a</b>–<b>d</b>) SiOC glass; (<b>e</b>–<b>h</b>) SiCaOC glass; (<b>i</b>–<b>l</b>) SiMgOC glass.</p>
Full article ">Figure 6
<p>SEM morphology and corresponding EDS spectra of glass materials after 14-day immersion in SBF: (<b>a</b>) SiCaOC and (<b>b</b>) SiMgOC.</p>
Full article ">Figure 7
<p>The change in different ions content and pH of SBF after the immersion tests for various glass materials: (<b>a</b>) [Si]; (<b>b</b>) [Ca]; (<b>c</b>) [P]; and (<b>d</b>) pH.</p>
Full article ">Figure 8
<p>Absorbance of the different samples in MTT toxicity test. n = 3, for each group (**** <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 9
<p>SEM images of L929 cells attached to the surface of the different samples: (<b>a</b>) SiOC; (<b>b</b>) SiCaOC; and (<b>c</b>) SiMgOC.</p>
Full article ">Figure 10
<p>Schematic illustration of the doping of Ca and Mg into the SiOC network structure.</p>
Full article ">
14 pages, 4034 KiB  
Article
In Situ Thermosensitive Mucoadhesive Nasal Gel Containing Sumatriptan: In Vitro and Ex Vivo Evaluations
by Aseel Alshraim, Doaa Alshora, Lubna Ashri, Ahlam Alhusaini, Nawal Alanazi and Nisreen M. Safwan
Polymers 2024, 16(23), 3422; https://doi.org/10.3390/polym16233422 - 5 Dec 2024
Viewed by 485
Abstract
The aim of this study was to develop a thermosensitive mucoadhesive (MA) in situ nasal gel for sumatriptan. A 3D response surface methodology (Design of Expert version 11) was employed to formulate nine different formulations. The Pluronic F-127 concentration (X1) and chitosan concentration [...] Read more.
The aim of this study was to develop a thermosensitive mucoadhesive (MA) in situ nasal gel for sumatriptan. A 3D response surface methodology (Design of Expert version 11) was employed to formulate nine different formulations. The Pluronic F-127 concentration (X1) and chitosan concentration (X2) were selected as independent factors. The formulas were studied in terms of pH, clarity, drug content, gelation temperature, gelation time, gel strength, MA strength, viscosity, % release after 5 h, and release kinetics. The optimized formulas were studied for % permeated after 5 h and stability in addition to previous tests. The study of the stability of the optimized formula was performed under accelerated conditions (40 ± 2 °C, 75 ± 5% RH) for 3 months. The outcomes of the optimized formula were a clear gel with a gelation temperature of 33 °C and a reasonable gelation time of less than one minute, and the release and permeation during 5 h were 40% and 50%, respectively. The formulated gel decreased the mucociliary clearance (MCC) and thus increased the retention time in the nasal cavity, resulting in enhancing SMT absorption, which could improve the drug efficacy. Full article
(This article belongs to the Special Issue Functional Gel and Their Multipurpose Applications)
Show Figures

Figure 1

Figure 1
<p>Effect of Pluronic F-127 on the gelation temperature (<b>a</b>), gelation strength (<b>b</b>), mucoadhesive strength (<b>c</b>), and % release after 5 h (<b>d</b>).</p>
Full article ">Figure 2
<p>Effect of chitosan HMW on the gelation temperature (<b>a</b>), gelation strength (<b>b</b>), mucoadhesive strength (<b>c</b>), and % release after 5 h (<b>d</b>).</p>
Full article ">Figure 3
<p>A 3D response plot for the gelation temperature (<b>a</b>), gelation strength (<b>b</b>), mucoadhesive strength (<b>c</b>), and % release after 5 h (<b>d</b>).</p>
Full article ">Figure 4
<p>SMT releases from the in situ nasal gel for each formula over 5 h.</p>
Full article ">Figure 5
<p>DSC thermograms for sumatriptan, chitosan, Pluromic F-127, and their physical mixture.</p>
Full article ">Figure 6
<p>FTIR spectra for SMT, chitosan, Plu F-127, and their physical mixture.</p>
Full article ">Figure 7
<p>Cumulative amount of sumatriptan permeated.</p>
Full article ">Figure 8
<p>Stability release study of sumatriptan from the in situ nasal gel optimized formula after 5 h at different conditions.</p>
Full article ">
24 pages, 18565 KiB  
Article
Injectable Photocrosslinked Hydrogel Dressing Encapsulating Quercetin-Loaded Zeolitic Imidazolate Framework-8 for Skin Wound Healing
by Zhao Chen, Man Zhe, Wenting Wu, Peiyun Yu, Yuzhen Xiao, Hao Liu, Ming Liu, Zhou Xiang and Fei Xing
Pharmaceutics 2024, 16(11), 1429; https://doi.org/10.3390/pharmaceutics16111429 - 10 Nov 2024
Viewed by 1054
Abstract
Background: Wound management is a critical component of clinical practice. Promoting timely healing of wounds is essential for patient recovery. Traditional treatments have limited efficacy due to prolonged healing times, excessive inflammatory responses, and susceptibility to infection. Methods: In this research, [...] Read more.
Background: Wound management is a critical component of clinical practice. Promoting timely healing of wounds is essential for patient recovery. Traditional treatments have limited efficacy due to prolonged healing times, excessive inflammatory responses, and susceptibility to infection. Methods: In this research, we created an injectable hydrogel wound dressing formulated from gelatin methacryloyl (GelMA) that encapsulates quercetin-loaded zeolitic imidazolate framework-8 (Qu@ZIF-8) nanoparticles. Next, its ability to promote skin wound healing was validated through in vitro experiments and animal studies. Results: Research conducted both in vitro and in vivo indicated that this hydrogel dressing effectively mitigates inflammation, inhibits bacterial growth, and promotes angiogenesis and collagen synthesis, thus facilitating a safe and efficient healing process for wounds. Conclusions: This cutting-edge scaffold system provides a novel strategy for wound repair and demonstrates significant potential for clinical applications. Full article
(This article belongs to the Topic New Nanomaterials for Biomedical Applications)
Show Figures

Figure 1

Figure 1
<p>Characterization of ZIF-8 and Qu@ZIF-8. (<b>A</b>) XRD patterns of Qu@ZIF-8, ZIF-8, and simulated ZIF-8. (<b>B</b>,<b>C</b>) SEM images and particle size distribution of ZIF-8 and Qu@ZIF-8. (<b>D</b>) XPS full spectrum of Qu@ZIF-8 and fine spectra of C 1s, Zn 2p and O 1s. (<b>E</b>) XPS full spectrum of ZIF-8 and XPS spectra of C 1s, Zn 2p and N 1s.</p>
Full article ">Figure 2
<p>Characterization of the GelMA and GelMA-based hydrogels. (<b>A</b>) <sup>1</sup>H-NMR spectra of gelatin and GelMA. a, b: acrylic protons of methacryloyl groups attached to lysine and hydroxylysine residues; c: methylene protons of lysine groups; d: methyl protons of methacryloyl groups. (<b>B</b>) TNBS assay for the degree of functionalization in GelMA. (<b>C</b>) Macroscopic images of the hydrogels before and after gelation. (<b>D</b>) Cryo-SEM images of the hydrogels at different magnifications.</p>
Full article ">Figure 3
<p>Characterization of the GelMA and GelMA-based hydrogels. (<b>A</b>) Quantitative assessment of the pore size within the hydrogels. (<b>B</b>) Quantitative analysis of the hydrogels’ porosity. (<b>C</b>) Rheological analysis of the hydrogels in a dynamic frequency sweep experiment. (<b>D</b>,<b>E</b>) Stress–strain curves and compression Young’s modulus of the hydrogels. (<b>F</b>) Swelling ratio–time curve of the hydrogels. (<b>G</b>) Degradation rate–time curve of the hydrogels. (<b>H</b>) Cumulative release rate–time curves of quercetin from the Qu-Gel and Qu@ZIF-8-Gel hydrogels in PBS (pH 5.6 and 7.4). (<b>I</b>) Cumulative release rate–time curves of zinc ions from the ZIF-8-Gel and Qu@ZIF-8-Gel hydrogels in PBS (pH 5.6 and 7.4). ns: not significant.</p>
Full article ">Figure 4
<p>Evaluation of the in vitro biocompatibility of the hydrogels and their drug components. (<b>A</b>) NIH-3T3 cells viability with different drug treatments. (<b>B</b>) Schematic diagram of the Transwell system utilized for co-culturing hydrogels with NIH-3T3 cells. (<b>C</b>,<b>F</b>) NIH-3T3 cell viability exposed to various hydrogels and its quantification. (<b>D</b>) NIH-3T3 cell morphology with different hydrogel treatments. (<b>E</b>,<b>G</b>) NIH-3T3 cell migration exposed to various hydrogels and its quantification. (<b>H</b>) NIH-3T3 cell viability exposed to various hydrogels. (<b>I</b>) NIH-3T3 cell adherence exposed to various hydrogels. * <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; ns: not significant.</p>
Full article ">Figure 5
<p>Evaluation of the immunomodulatory and pro-angiogenic effects of the hydrogels in vitro. (<b>A</b>) Quantification of IL-10 and IL-12 in RAW264.7 cell cultures with different treatments. (<b>B</b>) Quantification of iNOS and CD206 genes in RAW264.7 cells with different treatments. (<b>C</b>) Immunofluorescence with CD68, CD86, and CD206 in RAW264.7 cells with different treatments. (<b>D</b>) Immunofluorescence with VEGF in HUVEC cells with different treatments. (<b>E</b>) Quantification of HIF-1α and VEGF genes in HUVEC cells with different treatments. (<b>F</b>,<b>G</b>) HUVEC cell angiogenesis with different treatments and its quantification. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 5 Cont.
<p>Evaluation of the immunomodulatory and pro-angiogenic effects of the hydrogels in vitro. (<b>A</b>) Quantification of IL-10 and IL-12 in RAW264.7 cell cultures with different treatments. (<b>B</b>) Quantification of iNOS and CD206 genes in RAW264.7 cells with different treatments. (<b>C</b>) Immunofluorescence with CD68, CD86, and CD206 in RAW264.7 cells with different treatments. (<b>D</b>) Immunofluorescence with VEGF in HUVEC cells with different treatments. (<b>E</b>) Quantification of HIF-1α and VEGF genes in HUVEC cells with different treatments. (<b>F</b>,<b>G</b>) HUVEC cell angiogenesis with different treatments and its quantification. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Evaluation of the in vitro antibacterial efficacy of the hydrogels. (<b>A</b>) Schematic illustration of the antibacterial action of Qu@ZIF-8 via the release of quercetin and zinc ions. (<b>B</b>,<b>C</b>) Images of <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span> colonies on agar broth plates exposed to various hydrogels, along with quantitative analysis of antibacterial efficiency. (<b>D</b>) SEM images of <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span> exposed to various hydrogels. * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7
<p>In vivo assessment of wound healing. (<b>A</b>) Schematic representation of the construction and testing process of the animal model. (<b>B</b>,<b>C</b>) Macroscopic photographs depicting the wound-healing progression in rats subjected to various treatments, along with quantification of the wound-healing rates on days 4, 7, and 14. (<b>D</b>) H&amp;E staining of wound samples exposed to various treatments. (<b>E</b>) Immunofluorescence with CD31 and α-SMA in wound samples exposed to various treatments. (<b>F</b>,<b>G</b>) Immunohistochemical images along with quantitative analysis of CD86, CD206, Collagen I, and Collagen III in wound samples exposed to various treatments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 7 Cont.
<p>In vivo assessment of wound healing. (<b>A</b>) Schematic representation of the construction and testing process of the animal model. (<b>B</b>,<b>C</b>) Macroscopic photographs depicting the wound-healing progression in rats subjected to various treatments, along with quantification of the wound-healing rates on days 4, 7, and 14. (<b>D</b>) H&amp;E staining of wound samples exposed to various treatments. (<b>E</b>) Immunofluorescence with CD31 and α-SMA in wound samples exposed to various treatments. (<b>F</b>,<b>G</b>) Immunohistochemical images along with quantitative analysis of CD86, CD206, Collagen I, and Collagen III in wound samples exposed to various treatments. * <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 ">Scheme 1
<p>Preparation of the Qu@ZIF-8-Gel hydrogel and its function in the wound-healing process (Created with <a href="http://BioRender.com" target="_blank">BioRender.com</a>).</p>
Full article ">
12 pages, 2433 KiB  
Article
High-Performance Sol–Gel-Derived CNT-ZnO Nanocomposite-Based Photodetectors with Controlled Surface Wrinkles
by Hee-Jin Kim, Seung Hun Lee, Dabin Jeon and Sung-Nam Lee
Materials 2024, 17(21), 5325; https://doi.org/10.3390/ma17215325 - 31 Oct 2024
Viewed by 630
Abstract
We investigate the effects of incorporating single-walled carbon nanotubes (CNTs) into sol–gel-derived ZnO thin films to enhance their optoelectronic properties for photodetector applications. ZnO thin films were fabricated on c-plane sapphire substrates with varying CNT concentrations ranging from 0 to 2.0 wt%. Characterization [...] Read more.
We investigate the effects of incorporating single-walled carbon nanotubes (CNTs) into sol–gel-derived ZnO thin films to enhance their optoelectronic properties for photodetector applications. ZnO thin films were fabricated on c-plane sapphire substrates with varying CNT concentrations ranging from 0 to 2.0 wt%. Characterization techniques, including high-resolution X-ray diffraction, photoluminescence, and atomic force microscopy, demonstrated the preferential growth of the ZnO (002) facet and improved optical properties with the increase in the CNT content. Electrical measurements revealed that the optimal CNT concentration of 1.5 wt% resulted in a significant increase in the dark current (from 0.34 mA to 1.7 mA) and peak photocurrent (502.9 µA), along with enhanced photoresponsivity. The rising and falling times of the photocurrent were notably reduced at this concentration, indicating improved charge dynamics due to the formation of a p-CNT/n-ZnO heterojunction. The findings suggest that the incorporation of CNTs not only modifies the structural and optical characteristics of ZnO thin films but also significantly enhances their electrical performance, positioning CNT-ZnO composites as promising candidates for advanced photodetector technologies in optoelectronic applications. Full article
(This article belongs to the Special Issue Advanced and Smart Materials in Photoelectric Applications)
Show Figures

Figure 1

Figure 1
<p>AFM images showing the surface morphology of ZnO thin films incorporated with CNTs at concentrations of (<b>a</b>) 0%, (<b>b</b>) 0.5 wt%, (<b>c</b>) 1.0 wt%, (<b>d</b>) 1.5 wt%, and (<b>e</b>) 2.0 wt%, fabricated using the sol–gel method. (<b>f</b>) RMS roughness of sol–gel-derived ZnO thin films embedded with varying concentrations of CNTs. Insets are photographic images of ZnO/sapphire and 2.0 wt% CNT-ZnO/sapphire.</p>
Full article ">Figure 2
<p>High-resolution X-ray diffraction (HR-XRD) patterns of ZnO thin films with varying CNT concentrations (0 to 2.0 wt%) deposited on c-plane sapphire substrates using the sol–gel method.</p>
Full article ">Figure 3
<p>(<b>a</b>) Room temperature photoluminescence spectra of CNT-ZnO nanocomposite thin films excited with a 266 nm light source, and (<b>b</b>) absorbance spectra of CNT-ZnO nanocomposite thin films as a function of wavelength measured using ultraviolet–visible spectroscopy. Insets of (<b>a</b>,<b>b</b>) show the PL band-edge intensity, full width at half maximum (FWHM), and the (αhν)<sup>2</sup> (10<sup>11</sup> eV<sup>2</sup> cm<sup>−2</sup>) values for sol–gel-derived ZnO thin films embedded with varying CNT concentrations, presented as a function of CNT incorporation.</p>
Full article ">Figure 4
<p>(<b>a</b>) Carrier concentration and (<b>b</b>) resistivity as a function of CNT concentration in CNT-ZnO nanocomposite thin films, measured using the van der Pauw method.</p>
Full article ">Figure 5
<p>(<b>a</b>) Dark current–voltage and (<b>b</b>) photocurrent–voltage characteristics of the Al/ZnO-CNT/Al photodetector. The insets in (<b>a</b>,<b>b</b>) are a graph of the dark current as a function of CNT content at an applied voltage of 5.0 V and a schematic diagram of a CNT-ZnO nanocomposite photodetector with UV light applied, respectively. (<b>c</b>) Photocurrent and (<b>d</b>) photoresponsivity as a function of CNT content under a 5.0 V applied voltage.</p>
Full article ">Figure 6
<p>Photocurrent as a function of time after 950 s of excitation and termination of 365 nm ultraviolet light on an Al/ZnO-CNT/Al photodetector: (<b>a</b>) photocurrent response upon UV light application, (<b>b</b>) maximum photocurrent, (<b>c</b>) rise time, and (<b>d</b>) fall time after UV light termination as a function of the CNT content.</p>
Full article ">
10 pages, 2355 KiB  
Communication
Efficacy of Polyphenylene Carboxymethylene (PPCM) Gel at Protecting Type I Interferon Receptors Knockout Mice from Intravaginal Ebola Virus Challenge
by Olivier Escaffre, Terry L. Juelich, Jennifer K. Smith, Lihong Zhang, Madison Pearson, Nigel Bourne and Alexander N. Freiberg
Viruses 2024, 16(11), 1693; https://doi.org/10.3390/v16111693 - 30 Oct 2024
Viewed by 597
Abstract
Ebola virus (EBOV) is one of three filovirus members of the Orthoebolavirus genus that can cause severe Ebola disease (EBOD) in humans. Transmission predominantly occurs from spillover events from wildlife but has also happened between humans with infected bodily fluids. Specifically, the sexual [...] Read more.
Ebola virus (EBOV) is one of three filovirus members of the Orthoebolavirus genus that can cause severe Ebola disease (EBOD) in humans. Transmission predominantly occurs from spillover events from wildlife but has also happened between humans with infected bodily fluids. Specifically, the sexual route through infectious male survivors could be the origin of flare up events leading to the deaths of multiple women. More studies are needed to comprehend this route of infection which has recently received more focus. The use of microbicides prior to intercourse is of interest if neither of the Ebola vaccines are an option. These experimental products have been used against sexually transmitted diseases, and recently polyphenylene carboxymethylene (PPCM) showed efficacy against EBOV in vitro. Shortly after, the first animal model of EBOV sexual transmission was established using type I interferon receptors (IFNAR−/−) knockout female mice in which mortality endpoint could be achieved. Here, we investigated PPCM efficacy against a mouse-adapted (ma)EBOV isolate in IFNAR−/− mice and demonstrated that 4% PPCM gel caused a 20% reduction in mortality in two distinct groups compared to control groups when inoculated prior to virus challenge. Among animals that succumbed to disease despite PPCM treatment, we report an increase in median survival time as well as a less infectious virus, and fewer virus positive vaginal swabs compared to those from vehicle-treated animals, altogether indicating the beneficial effect of using PPCM prior to exposure. A post-study analysis of the different gel formulations tested indicated that buffering the gels would have prevented an increase in acidity seen only in vehicles, suggesting that PPCM antiviral efficacy against EBOV was suboptimal in our experimental set-up. These results are encouraging and warrant further studies using optimized stable formulations with the goal of providing additional safe protective countermeasures from sexual transmission of EBOV in humans. Full article
Show Figures

Figure 1

Figure 1
<p>Intravaginal maEBOV challenge in progesterone-primed IFNAR<sup>−/−</sup> mice and PPCM efficacy at preventing infection. Four groups of mice (n = 10/group) received a topical gel intravaginally (vehicle or PPCM) and were then challenged with a 10<sup>4</sup> pfu dose and monitored for 21 days for weight loss (<b>A</b>,<b>B</b>), and mortality (<b>C</b>) as well as virus dissemination when moribund (<b>D</b>). Open symbols (<b>D</b>) are values obtained by RT-qPCR and substitute the value of undetected virus from titrating by plaque assay the other half of the corresponding tissue. Seroconversion was evaluated at day 21 post-challenge and PRNT<sub>50</sub> was calculated when possible (<b>E</b>). Shedding of virus was assessed longitudinally using vaginal swabs starting on day 2 (<b>F</b>,<b>G</b>). Note that replicates in (<b>F</b>,<b>G</b>) are aligned for better tracking of individual values over time and some symbols may overlap in the early time points (n = 10 in all groups at days 2 and 4). Depending on the assay, the limit of detection was 10<sup>2</sup> PFU/gram (<b>D</b>, thick horizontal dotted line) or 10<sup>1.82</sup> PFU/mL (<b>F</b>,<b>G</b>) for titration by plaque assay, or 10<sup>0.5</sup> equivalent (equ) PFU/gram (<b>D</b>, thin horizontal dotted line) by RT-qPCR analysis with the virus stock used as standard. Subjects are represented by individual symbols in (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>), and may overlap in (<b>F</b>,<b>G</b>). The Mantel–Cox test was applied to survival data. The <span class="html-italic">t</span>-test was applied to titration data from swabs. Asterisks (* or **) indicate statistical differences of <span class="html-italic">p</span> &lt; 0.05 or 0.01, respectively, in the survival data (<b>C</b>) or average virus titers from swabs (<b>F</b>,<b>G</b>) between groups. (ns) for non-significant. Abbreviation: Genital tract (G.T.).</p>
Full article ">Figure 2
<p>Hematology and biochemical profile of blood from IFNAR<sup>−/−</sup> mice following intravaginal maEBOV challenge. Bar graphs represent lymphocyte (<b>A</b>) and neutrophil (<b>B</b>) populations from whole blood analyses as well as levels of globulin (<b>C</b>), glucose (<b>D</b>), and albumin (<b>E</b>) in sera at a predetermined time or when the subject became moribund. Subjects are represented by individual symbols and error bars show standard deviations. Note that not all parameters could be determined from a sample due to low volume or hemolysis. A one-way ANOVA, followed by Tukey’s multiple comparisons test, was used. Asterisks (*, **, ****) indicate statistical differences of <span class="html-italic">p</span> &lt; 0.05, 0.01, or 0.0001, respectively, for a given parameter between groups.</p>
Full article ">Figure 3
<p>Circulating inflammatory markers resulting from intravaginal maEBOV challenge in progesterone-primed IFNAR<sup>−/−</sup> mice. Serum samples were collected at a predetermined time or when the subject became moribund and were then assessed by multiplex immunoassays. Subjects are represented by individual symbols, and error bars show standard deviations. Note that not all samples were available due to low blood volumes collected from moribund animals. A one-way ANOVA, followed by Tukey’s multiple comparisons test, was used. Asterisks (*, **, ***, ****) indicate statistical differences of <span class="html-italic">p</span> &lt; 0.05, 0.01, 0.001, or 0.0001, respectively, for a given analyte between groups. (ns) for non-significant.</p>
Full article ">
23 pages, 6798 KiB  
Review
Advanced Hybrid Strategies of GelMA Composite Hydrogels in Bone Defect Repair
by Han Yu, Xi Luo, Yanling Li, Lei Shao, Fang Yang, Qian Pang, Yabin Zhu and Ruixia Hou
Polymers 2024, 16(21), 3039; https://doi.org/10.3390/polym16213039 - 29 Oct 2024
Viewed by 1418
Abstract
To date, severe bone defects remain a significant challenge to the quality of life. All clinically used bone grafts have their limitations. Bone tissue engineering offers the promise of novel bone graft substitutes. Various biomaterial scaffolds are fabricated by mimicking the natural bone [...] Read more.
To date, severe bone defects remain a significant challenge to the quality of life. All clinically used bone grafts have their limitations. Bone tissue engineering offers the promise of novel bone graft substitutes. Various biomaterial scaffolds are fabricated by mimicking the natural bone structure, mechanical properties, and biological properties. Among them, gelatin methacryloyl (GelMA), as a modified natural biomaterial, possesses a controllable chemical network, high cellular stability and viability, good biocompatibility and degradability, and holds the prospect of a wide range of applications. However, because they are hindered by their mechanical properties, degradation rate, and lack of osteogenic activity, GelMA hydrogels need to be combined with other materials to improve the properties of the composites and endow them with the ability for osteogenesis, vascularization, and neurogenesis. In this paper, we systematically review and summarize the research progress of GelMA composite hydrogel scaffolds in the field of bone defect repair, and discuss ways to improve the properties, which will provide ideas for the design and application of bionic bone substitutes. Full article
(This article belongs to the Special Issue Bioactive and Biomedical Hydrogel Dressings for Wound Healing)
Show Figures

Figure 1

Figure 1
<p>Synthesis of gelatin methacryloyl (GelMA) hydrogel and its properties. (<b>A</b>) Scheme for the preparation of photocrosslinked GelMA hydrogel [<a href="#B18-polymers-16-03039" class="html-bibr">18</a>]. Copyright © 2015 Elsevier; (<b>B</b>) (<b>i</b>) 1 HNMR spectra of unmodified gelatin and GelMA with different DMS (dimethyl sulfide) values. The left and right red rectangles represent the signals of aromatic amino acid and lysine amino acid, respectively; (<b>ii</b>) the relationship between DMS of GelMA and the MA–gelatin feed ratio (* <span class="html-italic">p</span> &lt; 0.05) [<a href="#B20-polymers-16-03039" class="html-bibr">20</a>]. Copyright © 2015 Springer Nature; (<b>C</b>) scanning electron microscopy (SEM) images of homologous series of GelMA hydrogels showing the influence of hydrogel %DOF, monomer concentration, and sample geometry: (<b>a</b> vs. <b>b</b>) 10 vs. 7 wt%, 35 % DOF, 1mm thick. (<b>b</b> vs. <b>c</b>) 1 vs. 2 mm thick, 10 wt%, 35 % DOF. (<b>a</b> vs. <b>d</b>) 35 vs. 85 % DOF, 10 wt%, 1 mm thick. (<b>d</b> vs. <b>e</b>) 10 vs. 7 wt%, 85 % DOF, 1 mm thick. (<b>e</b> vs. <b>f</b>) 2 mm thick photopolymerized in Teflon mold vs. between glass slides. Scale bars: 100 μm [<a href="#B24-polymers-16-03039" class="html-bibr">24</a>]. Copyright © 2013 Wiley; (<b>D</b>) attachment of MG63 osteoblasts on glass and GelMA [<a href="#B18-polymers-16-03039" class="html-bibr">18</a>,<a href="#B25-polymers-16-03039" class="html-bibr">25</a>], copyright © 2017 IOP Publishing.</p>
Full article ">Figure 2
<p>Enhanced mechanical performance. (<b>A</b>) Compressive modulus of hydrogels made of GelMA with various DM [<a href="#B34-polymers-16-03039" class="html-bibr">34</a>]. Copyright © 2012 Wiley; (<b>B</b>) compressive modulus of hydrogels made of different concentrations of GelMA [<a href="#B35-polymers-16-03039" class="html-bibr">35</a>]. Copyright © 2013 Wiley; (<b>C</b>) effect of UV exposure time (10–120 min) on the temperature dependence of the elastic modulus [<a href="#B23-polymers-16-03039" class="html-bibr">23</a>], copyright © 2000 American Chemical Society; (<b>D</b>) effect of LAP concentration on G’-plateau; (<b>E</b>) effect of the laponite in different ink on the mechanical strength; (<b>F</b>) time sweep of storage and loss moduli (G’/G’’) [<a href="#B37-polymers-16-03039" class="html-bibr">37</a>]; (<b>G</b>) Young’s modulus of hybrid hydrogels of 5 wt% GelMA and different concentrations of submicron line [<a href="#B38-polymers-16-03039" class="html-bibr">38</a>]. Copyright © 2021 Elsevier: (<b>H</b>) Mechanical characterization of acellular printed gel; (<b>I</b>) Effect on the hydrogels with 10 wt% GelMA and varied initial concentrations of PACG [<a href="#B38-polymers-16-03039" class="html-bibr">38</a>]. Copyright © 2021 IOP Publishing; (ns ≙ <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 3
<p>Mitigating the degradation of GelMA-based hydrogels. (<b>A</b>) Equilibrium swelling properties of the GelMA hydrogels; (<b>B</b>) degradation profiles of hydrogels made with different concentrations of GelMA in PBS solution with collagenase at 37 °C [<a href="#B47-polymers-16-03039" class="html-bibr">47</a>], © 2020 by the authors; (<b>C</b>) comparison of the effects of Irgacure 2959 and LAP on the degradation properties of GelMA [<a href="#B48-polymers-16-03039" class="html-bibr">48</a>], copyright © 2020 IOP Publishing; (<b>D</b>) (i) synthesis of PEG–GelMA composite; (ii) degradation profiles of 5%wt PEGDMA and various ratios of GelMA in DPBS solution with collagenase; (iii) degradation profiles of 20%wt PEGDMA and various ratios of GelMA in DPBS solution with collagenase [<a href="#B49-polymers-16-03039" class="html-bibr">49</a>]. Copyright © 2022 Elsevier; (<b>E</b>) (i) the degradation behaviors of GelMA-H-nHA hydrogels; (ii) degradation behaviors of P(ACG-GelMA-L)-Mg2<sup>+</sup> hydrogels [<a href="#B50-polymers-16-03039" class="html-bibr">50</a>]. Copyright © 2022 Springer Nature; (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 4
<p>Endowing the hydrogel with osteogenic function with the addition of inorganic mineral materials. (<b>A</b>) SEM of GelMA/HAp composite scaffolds; (<b>B</b>) ALP staining of GelMA/HAp composite scaffolds [<a href="#B80-polymers-16-03039" class="html-bibr">80</a>], copyright © 2022 Elsevier; (<b>C</b>) digital images of GelMA-B hydrogels. (<b>D</b>) micro-CT analysis after implantation of GelMA-B; (<b>E</b>) quantitative analysis of regenerated bone volume with GelMA-B; (<b>F</b>) bone volume fraction in total tissue volume with GelMA-B [<a href="#B59-polymers-16-03039" class="html-bibr">59</a>]. Copyright © 2019 Elsevier; (<b>G</b>) SEM images: (i) GelMA; (ii) CNP (50%) GelMA; (iii) CNP GelMA; (<b>H</b>) (a) ALP activity of CNP GelMA; (b) gene expression study of CNP GelMA [<a href="#B81-polymers-16-03039" class="html-bibr">81</a>]; (Each group was compared with the control group: ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001; Comparison between GelMA group and GELMA-BH group: <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05,<sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 5
<p>Captures the release of metal ions to promote bone defect repair. (<b>A</b>) Scanning electron microscope images of GelMA-BP-Mg microspheres; (<b>B</b>) release curves of Mg<sup>2</sup>⁺ and BP from the composite microspheres; (<b>C</b>) Western blot of the GelMA-BP-Mg group; (<b>D</b>) Q-PCR of GelMA-BP-Mg group [<a href="#B91-polymers-16-03039" class="html-bibr">91</a>], copyright © 2021 American Chemistry Society; (<b>E</b>) X-ray test of M-Li.; (<b>F</b>) Li+ ions released from GM/M-Li hydrogel; (<b>G</b>) microfilament skeleton of BMSCs cultured in GM/M-Li hydrogels; (<b>H</b>) Western blotting of OCN, Runx2, and osterix proteins; (<b>I</b>) quantitative analysis of Runx2 and osterix proteins [<a href="#B92-polymers-16-03039" class="html-bibr">92</a>], copyright © 2022 Wiley; (ns ≙ <span class="html-italic">p</span> &gt; 0.05, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Adding signaling molecules, exosomes, and encapsulated seed cells to repair bone defects. (<b>A</b>) Photograph of GelMA-c-OGP solution and GelMA-c-OGP hydrogel; (<b>B</b>) the osteogenesis properties of GelMA-c-OGP hydrogel in vitro; (<b>C</b>) fabrication of GelMA-SN-SDF-1α hydrogel; (<b>D</b>) the area analyzed by micro-CT [<a href="#B72-polymers-16-03039" class="html-bibr">72</a>]. Copyright © 2019 Wiley; (<b>E</b>) electron microscope image of the engineering exosomes; (<b>F</b>) internalization of DiI-labelled EXOs into BMSCs; (<b>G</b>) exosome-expressed Bmp2 genes; (<b>H</b>) ALP expression and activity were assessed on days 3 and 7 [<a href="#B97-polymers-16-03039" class="html-bibr">97</a>]; (* <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Hybrid BP @ Mg or rGO stimulate nerve repair in bone. (<b>A</b>) GelMA-BP@Mg scaffolds; TEM assessment of BP nanosheets; TEM assessment of BP@Mg nanosheets; (<b>B</b>) stimulation of neural differentiation in cells on the GelMA, GelMA-BP, and GelMA-BP@Mg hydrogels; (<b>C</b>) relative neural-specific gene expression in cells on GelMA-BP@Mg hydrogel scaffolds (*, # Represents marked change in relation to GelMA and GelMA-BP hydrogel, respectively); (<b>D</b>) Typical confocal images of CD31 and β3-tubulin co-staining on specimen from the calvarial critical-sized defects [<a href="#B118-polymers-16-03039" class="html-bibr">118</a>]; (<b>E</b>) macroscopic and scanning electron microscope images of GelMA 3D scaffolds with different rGO concentrations; (<b>F</b>) myelination-related assessment of cultured Schwann cells; (<b>G</b>) H&amp;E, Masson staining, immunohistochemical staining, and immunofluorescence staining of tissue sections [<a href="#B119-polymers-16-03039" class="html-bibr">119</a>]. Copyright © 2013 Royal Society of Chemistry; (* <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; 0.0001).</p>
Full article ">Figure 8
<p>Bionic structure or adding DFO to accelerate bone vascular reconstruction and promote bone defect repair. (<b>A</b>) (i) Outline of construct design of microchanneled hydrogels; (ii) microscopic images and live–dead staining; (<b>B</b>) H&amp;E and positive TRAP staining after 8 weeks in vivo, revealing areas of red blood cell activity [<a href="#B126-polymers-16-03039" class="html-bibr">126</a>]. Copyright © 2018 Elsevier; (<b>C</b>) (i) representative 3D reconstructions of vessel formation; (ii) quantification of total vessel volume, degree of vessel anisotropy, vessel connectivity, and mean vessel thickness for all groups; (<b>D</b>) mechanistic insight into the Eth-DFO@GelMA/GGMA scaffold-induced osteogenesis and angiogenesis; (<b>E</b>) (a) eNOS expression; (b) HIF1-α expression; (c) SDF1-α expression after 3 and 7 days of culture of HUVECs in each group; (d) ELISA assay for VEGF content measurement after 3 days; (<b>F</b>) DFO release kinetics from DFO@GelMA/GGMA and Eth-DFO@GelMA/GGMA hydrogels’; (<b>G</b>) (i) quantification of the difference in expression of the associated proteins; (ii) immunofluorescence staining of CD31 (red), α-SMA (green), and HIF1-α (red) expression at 8 weeks [<a href="#B127-polymers-16-03039" class="html-bibr">127</a>]. Copyright © 2022 Elsevier; (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Scheme 1
<p>Schematic description of GelMA hydrogels hybridized with different materials, improve their shortcomings and endowing them with the functions of osseointegration, vascularization, and neuralization [<a href="#B18-polymers-16-03039" class="html-bibr">18</a>].</p>
Full article ">
18 pages, 7715 KiB  
Article
Pristine Photopolymerizable Gelatin Hydrogels: A Low-Cost and Easily Modifiable Platform for Biomedical Applications
by Maria Pérez-Araluce, Alessandro Cianciosi, Olalla Iglesias-García, Tomasz Jüngst, Carmen Sanmartín, Íñigo Navarro-Blasco, Felipe Prósper, Daniel Plano and Manuel M. Mazo
Antioxidants 2024, 13(10), 1238; https://doi.org/10.3390/antiox13101238 - 15 Oct 2024
Viewed by 1025
Abstract
The study addresses the challenge of temperature sensitivity in pristine gelatin hydrogels, widely used in biomedical applications due to their biocompatibility, low cost, and cell adhesion properties. Traditional gelatin hydrogels dissolve at physiological temperatures, limiting their utility. Here, we introduce a novel method [...] Read more.
The study addresses the challenge of temperature sensitivity in pristine gelatin hydrogels, widely used in biomedical applications due to their biocompatibility, low cost, and cell adhesion properties. Traditional gelatin hydrogels dissolve at physiological temperatures, limiting their utility. Here, we introduce a novel method for creating stable hydrogels at 37 °C using pristine gelatin through photopolymerization without requiring chemical modifications. This approach enhances consistency and simplifies production and functionalization of the gelatin with bioactive molecules. The stabilization mechanism involves the partial retention of the triple-helix structure of gelatin below 25 °C, which provides specific crosslinking sites. Upon activation by visible light, ruthenium (Ru) acts as a photosensitizer that generates sulphate radicals from sodium persulphate (SPS), inducing covalent bonding between tyrosine residues and “locking” the triple-helix conformation. The primary focus of this work is the characterization of the mechanical properties, swelling ratio, and biocompatibility of the photopolymerized gelatin hydrogels. Notably, these hydrogels supported better cell viability and elongation in normal human dermal fibroblasts (NHDFs) compared to GelMA, and similar performance was observed for human pluripotent stem cell-derived cardiomyocytes (hiPSC-CMs). As a proof of concept for functionalization, gelatin was modified with selenous acid (GelSe), which demonstrated antioxidant and antimicrobial capacities, particularly against E. coli and S. aureus. These results suggest that pristine gelatin hydrogels, enhanced through this new photopolymerization method and functionalized with bioactive molecules, hold potential for advancing regenerative medicine and tissue engineering by providing robust, biocompatible scaffolds for cell culture and therapeutic applications. Full article
(This article belongs to the Special Issue Applications and Health Benefits of Novel Antioxidant Biomaterials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Hydrogel formation. Schematic representation of the polymerization process of pristine gelatin.</p>
Full article ">Figure 2
<p>Gelatin functionalization with selenium (GelSe). (<b>A</b>) Synthesis reaction of GelSe. (<b>B</b>) Schematic representation of GelSe hydrogel formation. (<b>C</b>) Schematic representation of the obtention of gelatin with different degree of functionalization (DoF) by mixing GelSe/GelSe-H (high DoF) with pristine gelatin in proportions 1:1 and 1:2 to obtain GelSe-M (medium DoF) and GelSe-L (low DoF), respectively.</p>
Full article ">Figure 3
<p>Rheological and photorheological tests (n = 3, 10 rad/s oscillation frequency, 10% shear strain) for the analysis of the viscosity and the viscoelastic properties of the pristine gelatin and polymerize pristine gelatin. (<b>A</b>) Viscosity profile of pristine gelatin at different concentrations (10, 7.5 and 5% <span class="html-italic">w</span>/<span class="html-italic">v</span>). (<b>B</b>) Temperature sweep of pristine gelatin. (<b>C</b>) Time sweep of gelatin at 37 °C (light activated at 20 s, 35 s exposure time, 7.5 cm light probe-to-sample distance). (<b>D</b>) Time sweep of gelatin previously incubated for 5 min at 21 °C (light activated at 20 s, 35 s exposure time, 7.5 cm light probe-to-sample distance). (<b>E</b>) Amplitude sweep of pristine gelatin hydrogel. (<b>F</b>) Frequency sweep of pristine gelatine hydrogel.</p>
Full article ">Figure 4
<p>Autofluorescence of dityrosine groups in response to UV light exposure. (<b>A</b>) Light-activated pristine gelatin hydrogels. (<b>B</b>) Non-activated pristine gelatin hydrogels. (<b>C</b>) Ru/SPS 1/10 mM solution in PBS. (<b>D</b>) Pristine gelatin 10% <span class="html-italic">w</span>/<span class="html-italic">v</span> in PBS.</p>
Full article ">Figure 5
<p>Swelling ratio. (<b>A</b>) Representation of the hydrogel diameter in cm after incubation in PBS at 37 °C over 30 days (N = 12). (<b>B</b>) Images showing the evolution of the hydrogel incubated in PBS at 37 °C over 30 days.</p>
Full article ">Figure 6
<p>Material addition to cell culture. (<b>A</b>) Alamar Blue assay of pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span> addition to NHDFs cultured in a 96-well plate. (<b>B</b>) Alamar Blue assay of pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span> addition to hiPSC-CMs cultured in a 96-well plate. (N = 3, unpaired <span class="html-italic">t</span>-test, ** <span class="html-italic">p</span> &lt; 0.005, ns: no significant differences).</p>
Full article ">Figure 7
<p>Cell encapsulation. (<b>A</b>) Alamar Blue assay of NHDFs encapsulated within pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span>. (<b>B</b>) Alamar Blue assay of hiPSC-CMs encapsulated within pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span>. (N = 3, unpaired <span class="html-italic">t</span>-test, * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.005, ns = no significative differences). (<b>C</b>) Fluorescence images of Live/Dead<sup>®</sup> assay of NHDFs encapsulated within pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span>. (<b>D</b>) Fluorescence images of Live/Dead assay of hiPSC-CMs encapsulated within pristine gelatin and GelMA 10% <span class="html-italic">w</span>/<span class="html-italic">v</span>. Live cells produced green fluorescence and dead cells showed red fluorescence.</p>
Full article ">Figure 8
<p>(<b>A</b>) <sup>77</sup><span class="html-italic">Se</span>-NMR of GelSe. (<b>B</b>) <sup>1</sup><span class="html-italic">H</span>-NMR of pristine gelatin (blue) and GelSe (red).</p>
Full article ">Figure 9
<p>Antioxidant properties of GelSe. DPPH analysis of GelSe-H. Results were calculated relative to a positive control (ascorbic acid 2 mg/mL). The “Ascorbic acid” result is the antioxidant capacity of ascorbic acid at 0.023 mg/mL, the equivalent concentration of selenium found in GelSe-H. (N = 3, unpaired <span class="html-italic">t</span>-test, *** <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 10
<p>Antibacterial properties of GelSe against <span class="html-italic">Escherichia coli</span> (<span class="html-italic">E. coli</span>) and <span class="html-italic">Staphylococcus aureus</span> (<span class="html-italic">S. aureus</span>). (<b>A</b>) Antibacterial properties of GelSe/GelSe-H. Agar plate culture of a sample from a 96-well plate containing 1 million bacteria of <span class="html-italic">E. coli</span> (left) and <span class="html-italic">S. aureus</span> (right), treated with five different concentrations of GelSe (GelSe-H.1 (13%), H.2 (10%), H.3.3 (6.67%), H.4 (5%), and H.5 (3.33% <span class="html-italic">w</span>/<span class="html-italic">v</span>)). (<b>B</b>) Antibacterial properties of GelSe-M.1 to GelSe-M.5 with <span class="html-italic">E. coli</span> (left) and <span class="html-italic">S. aureus</span> (right). (<b>C</b>) Antibacterial properties of GelSe-L.1 to GelSe-L.5 with <span class="html-italic">E. coli</span> (left) and <span class="html-italic">S. aureus</span> (right), including a control for each bacterial strain using pristine gelatin.</p>
Full article ">Figure 11
<p>Biocompatibility properties of different concentrations of GelSe-H and GelSe-M added to NHDFs. (<b>A</b>) Alamar blue assay (N = 3, unpaired <span class="html-italic">t</span>-test, *** <span class="html-italic">p</span> &lt; 0.001, ns = no significative differences). (<b>B</b>) Fluorescence images of Live/Dead<sup>®</sup> assay. Live cells produced green fluorescence and dead cells showed red fluorescence. Scale bars are 100 µm.</p>
Full article ">Figure 12
<p>Comparative table of GelSe-H and GelSe-M showing their antibacterial and biocompatibility properties. GelSe-M.3 is proposed as the best candidate, as it demonstrates good cell viability and antibacterial properties. X: not bactericidal/not biocompatible; ✓: bactericidal/biocompatible.</p>
Full article ">
18 pages, 4283 KiB  
Article
A Machine Learning Assisted Non-Enzymatic Electrochemical Biosensor to Detect Urea Based on Multi-Walled Carbon Nanotube Functionalized with Copper Oxide Micro-Flowers
by Jitendra B. Zalke, Manish L. Bhaiyya, Pooja A. Jain, Devashree N. Sakharkar, Jayu Kalambe, Nitin P. Narkhede, Mangesh B. Thakre, Dinesh R. Rotake, Madhusudan B. Kulkarni and Shiv Govind Singh
Biosensors 2024, 14(10), 504; https://doi.org/10.3390/bios14100504 - 15 Oct 2024
Viewed by 1340
Abstract
Detecting urea is crucial for diagnosing related health conditions and ensuring timely medical intervention. The addition of machine learning (ML) technologies has completely changed the field of biochemical sensing, providing enhanced accuracy and reliability. In the present work, an ML-assisted screen-printed, flexible, electrochemical, [...] Read more.
Detecting urea is crucial for diagnosing related health conditions and ensuring timely medical intervention. The addition of machine learning (ML) technologies has completely changed the field of biochemical sensing, providing enhanced accuracy and reliability. In the present work, an ML-assisted screen-printed, flexible, electrochemical, non-enzymatic biosensor was proposed to quantify urea concentrations. For the detection of urea, the biosensor was modified with a multi-walled carbon nanotube-zinc oxide (MWCNT-ZnO) nanocomposite functionalized with copper oxide (CuO) micro-flowers (MFs). Further, the CuO-MFs were synthesized using a standard sol-gel approach, and the obtained particles were subjected to various characterization techniques, including X-ray diffraction (XRD), field emission scanning electron microscopy (FESEM), and Fourier transform infrared (FTIR) spectroscopy. The sensor’s performance for urea detection was evaluated by assessing the dependence of peak currents on analyte concentration using cyclic voltammetry (CV) at different scan rates of 50, 75, and 100 mV/s. The designed non-enzymatic biosensor showed an acceptable linear range of operation of 0.5–8 mM, and the limit of detection (LoD) observed was 78.479 nM, which is well aligned with the urea concentration found in human blood and exhibits a good sensitivity of 117.98 mA mM−1 cm−2. Additionally, different regression-based ML models were applied to determine CV parameters to predict urea concentrations experimentally. ML significantly improves the accuracy and reliability of screen-printed biosensors, enabling accurate predictions of urea levels. Finally, the combination of ML and biosensor design emphasizes not only the high sensitivity and accuracy of the sensor but also its potential for complex non-enzymatic urea detection applications. Future advancements in accurate biochemical sensing technologies are made possible by this strong and dependable methodology. Full article
(This article belongs to the Special Issue Advances in Biosensing and Bioanalysis Based on Nanozymes)
Show Figures

Figure 1

Figure 1
<p>Process steps to detect urea, based on MWCNT-ZnO functionalized with novel CuO-MF and ML-approach.</p>
Full article ">Figure 2
<p>(<b>A</b>) Process steps for synthesis and preparation of CuO-MFs. (<b>B</b>) Functionalization of Gii-Sens Integrated Graphene SPE.</p>
Full article ">Figure 3
<p>(<b>A</b>) SEM image of MWCNT−ZnO showing the exact morphology of nanofibers, (<b>B</b>) EDX spectrum of MWCNT−ZnO nanofibers showing the elemental content material composition, (<b>C</b>) XRD analysis provides the crystalline structure of the synthesized MWCNT−ZnO nanofibers, (<b>D</b>) FTIR analysis showing the functional groups in the MWCNT−ZnO composite.</p>
Full article ">Figure 4
<p>(<b>A</b>) 3D pictures of the CuO-MFs, derived from a morphological study performed using scanning electron microscopy (SEM). (<b>B</b>) EDX spectra of CuO-MFs. (<b>C</b>) XRD spectra of CuO-MFs. (<b>D</b>) FTIR spectrum of CuO-MFs.</p>
Full article ">Figure 5
<p>SEM image of MWCNT-ZnO/CuO micro-flower deposition on working electrode of Integrated Graphene IG-GII-SENS-01 SPE at magnification levels of (<b>A</b>) 1 mm, (<b>B</b>) 10 µm, (<b>C</b>) 1 µm, and (<b>D</b>) 100 nm.</p>
Full article ">Figure 6
<p>Cyclic voltammetry (CV) responses of the MWCNT−ZnO/CuO−MFs modified SPEs were measured at various concentrations of urea ranging from 0.5 to 10 mM in the presence of a 5 mM solution of Ferroferricyanide [Fe(CN)<sub>6</sub>]<sup>3−4−</sup> as the standard redox probe at scan rates of (<b>A</b>) 50 mV/s, (<b>B</b>) 75 mV/s, and (<b>C</b>) 100 mV/s. Corresponding calibration plot of urea concentration (mM) versus current (A/cm<sup>2</sup>) for a scan rate of (<b>D</b>) 50 mV/s, (<b>E</b>) 75 mV/s, (<b>F</b>) 100 mV/s (n = 5).</p>
Full article ">Figure 7
<p>(<b>A</b>) Selectivity study with commonly identified interfering elements in human blood such as Galactose, Dextrose, Maltose, Lactose, Ascorbic Acid, and Uric Acid with a concentration of 0.1 mM at the scan rate of 100 mV/se, (<b>B</b>) Stability study of MWCNT−ZnO/CuO−MFs modified sensor for urea detection.</p>
Full article ">Figure 8
<p>Prediction of urea concentration using machine learning algorithms: The available machine learning models are (<b>A</b>) LR, (<b>B</b>) DT, (<b>C</b>) RF, (<b>D</b>) KNN, (<b>E</b>) AdaBoost, and (<b>F</b>) GB.</p>
Full article ">
17 pages, 6452 KiB  
Article
3D-Printable Gelatin Methacrylate-Xanthan Gum Hydrogel Bioink Enabling Human Induced Pluripotent Stem Cell Differentiation into Cardiomyocytes
by Virginia Deidda, Isabel Ventisette, Marianna Langione, Lucrezia Giammarino, Josè Manuel Pioner, Caterina Credi and Federico Carpi
J. Funct. Biomater. 2024, 15(10), 297; https://doi.org/10.3390/jfb15100297 - 5 Oct 2024
Viewed by 1285
Abstract
We describe the development of a bioink to bioprint human induced pluripotent stem cells (hiPSCs) for possible cardiac tissue engineering using a gelatin methacrylate (GelMA)-based hydrogel. While previous studies have shown that GelMA at a low concentration (5% w/v) allows [...] Read more.
We describe the development of a bioink to bioprint human induced pluripotent stem cells (hiPSCs) for possible cardiac tissue engineering using a gelatin methacrylate (GelMA)-based hydrogel. While previous studies have shown that GelMA at a low concentration (5% w/v) allows for the growth of diverse cells, its 3D printability has been found to be limited by its low viscosity. To overcome that drawback, making the hydrogel both compatible with hiPSCs and 3D-printable, we developed an extrudable GelMA-based bioink by adding xanthan gum (XG). The GelMA-XG composite hydrogel had an elastic modulus (~9 kPa) comparable to that of cardiac tissue, and enabled 3D printing with high values of printing accuracy (83%) and printability (0.98). Tests with hiPSCs showed the hydrogel’s ability to promote their proliferation within both 2D and 3D cell cultures. The tests also showed that hiPSCs inside hemispheres of the hydrogel were able to differentiate into cardiomyocytes, capable of spontaneous contractions (average frequency of ~0.5 Hz and amplitude of ~2%). Furthermore, bioprinting tests proved the possibility of fabricating 3D constructs of the hiPSC-laden hydrogel, with well-defined line widths (~800 μm). Full article
Show Figures

Figure 1

Figure 1
<p>Swelling ratio as a function of time for the GelMA and GelMA-XG hydrogels. The error bars represent the standard deviation among four samples. A Wilcoxon–Mann–Whitney test indicated a statistically non-significant difference between the GelMA and GelMA-XG series of data (<span class="html-italic">p</span> = 0.2975).</p>
Full article ">Figure 2
<p>Stress-relative indentation curves of the GelMA and GelMA-XG hydrogels. The error bars represent the standard deviation among four samples. A Wilcoxon–Mann–Whitney test indicated a statistically significant difference between the GelMA and GelMA-XG series of data (<span class="html-italic">p</span> = 0.0001942).</p>
Full article ">Figure 3
<p>Degradation ratio as a function of time for the GelMA and GelMA-XG hydrogels. The error bars represent the standard deviation among two samples. A Wilcoxon–Mann–Whitney test indicated a statistically non-significant difference between the GelMA and GelMA-XG series of data (<span class="html-italic">p</span> = 0.3624).</p>
Full article ">Figure 4
<p>Example of a structure 3D printed with the GelMA-XG hydrogel.</p>
Full article ">Figure 5
<p>Assessment of the quality of the extrudable constructs made of the GelMA-XG hydrogel: (<b>A</b>) square frame and example of a 3D-printed version, used to quantify the printing accuracy; (<b>B</b>) grid of square pores and example of a 3D-printed version, used to quantify the printability; zoomed-in views on the right-hand side panel show detailed areas (1, 2, 3) of the printed grid.</p>
Full article ">Figure 6
<p>Optical microscopy images of hiPSCs within 2D cell cultures on top of layers of the GelMA-XG, GelMA, GelMA-XG-FN and GelMA-FN hydrogels, taken after 4 and 9 days from seeding.</p>
Full article ">Figure 7
<p>Optical microscopy images of hiPSCs within 3D cell cultures inside hydrogel hemispheres made of GelMA-XG and GelMA-XG-FN. The images were taken after 2 and 7 days from UV cross-linking for 15 s, which occurred above a substrate maintained at a temperature of 25 °C or 15 °C.</p>
Full article ">Figure 8
<p>Fluorescence microscopy images of the GelMA-XG hydrogel hemispheres containing differentiated hiPSC cardiomyocytes at day 8 of the differentiation process. The differentiation is indicated by the endogenous expression of the mEGFP fluorescent marker. See the videos in the <a href="#app1-jfb-15-00297" class="html-app">Supplementary Materials</a> which show specimens contracting at a frequency of ~0.8 Hz (highest value recorded).</p>
Full article ">Figure 9
<p>Action potential duration at 90% repolarisation of the differentiated cardiomyocytes inside the GelMA-XG hydrogel hemispheres, on day 17 of the differentiation process, in response to an electrical stimulation at 0.5 Hz or 1 Hz, at 37 °C. The error bars represent the standard deviation among three samples.</p>
Full article ">Figure 10
<p>Optical microscopy images of 3D-printed structures made of the hiPSC-laden GelMA-XG hydrogel.</p>
Full article ">
15 pages, 3627 KiB  
Article
Photo-Crosslinked Polyurethane—Containing Gel Polymer Electrolytes via Free-Radical Polymerization Method
by Fatmanur Uyumaz, Yerkezhan Yerkinbekova, Sandugash Kalybekkyzy and Memet Vezir Kahraman
Polymers 2024, 16(18), 2628; https://doi.org/10.3390/polym16182628 - 18 Sep 2024
Viewed by 1239
Abstract
Using a novel technique, crosslinked gel polymer electrolytes (GPEs) designed for lithium-ion battery applications have been created. To form the photo crosslink via free-radical polymerization, a mixture of polyurethane acrylate (PUA), polyurethane methacrylate (PUMA), vinyl phosphonic acid (VPA), and bis[2-(methacryloyloxy)ethyl] phosphate (BMEP) was [...] Read more.
Using a novel technique, crosslinked gel polymer electrolytes (GPEs) designed for lithium-ion battery applications have been created. To form the photo crosslink via free-radical polymerization, a mixture of polyurethane acrylate (PUA), polyurethane methacrylate (PUMA), vinyl phosphonic acid (VPA), and bis[2-(methacryloyloxy)ethyl] phosphate (BMEP) was exposed to ultraviolet (UV) radiation during the fabrication process. The unique crosslinked configuration of the membrane increased its stability and made it suitable for use with liquid electrolytes. The resulting GPE has a much higher ionic conductivity (1.83 × 10−3 S cm−1) than the commercially available Celgrad2500 separator. A crosslinked structure formed by the hydrophilic properties of the PUA-PUMA blend and the higher phosphate content from BMEP reduced the leakage of the electrolyte solution while at the same time providing a greater capacity for liquid retention, significantly improving the mechanical and thermal stability of the membrane. GPP2 shows electrochemical stability up to 3.78 V. The coin cell that was assembled with a LiFePO4 cathode had remarkable cycling characteristics and generated a high reversible capacity of 149 mA h g−1 at 0.1 C. It also managed to maintain a consistent Coulombic efficiency of almost 100%. Furthermore, 91.5% of the original discharge capacity was maintained. However, the improved ionic conductivity, superior electrochemical performance, and high safety of GPEs hold great promise for the development of flexible energy storage systems in the future. Full article
(This article belongs to the Section Polymer Membranes and Films)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The PP2 membrane’s FTIR spectrum and (<b>b</b>) the PP2 membrane’s composition and photo illustration.</p>
Full article ">Figure 2
<p>Schematic representation of the synthesis of the UV-cured crosslinked membrane PUA/PUMA/BMEP (GPP2) gel polymer electrolyte.</p>
Full article ">Figure 3
<p>SEM images of the membranes: (<b>a</b>) PP1; (<b>b</b>) PP2; (<b>c</b>) PP3.</p>
Full article ">Figure 4
<p>(<b>a</b>) Thermogravimetric analysis (TGA), and (<b>b</b>) Differential Scanning Calorimetry results of membranes.</p>
Full article ">Figure 5
<p>The following images show (<b>a</b>) liquid electrolyte uptake charts and (<b>b</b>) liquid electrolyte leakage test results.</p>
Full article ">Figure 6
<p>(<b>a</b>) The outcomes of electrochemical impedance spectroscopy on symmetric stainless-steel (SS) electrodes using GPP membranes. (<b>b</b>) The linear sweep voltammogram for the Li/GPP2/SS cell. (<b>c</b>) The initial charge–discharge profiles of GPP2. (<b>d</b>) GPP2 rate performance. (<b>e</b>) Galvanostatic cyclability and Coulombic efficiency of Li/GPP2/LiFePO<sub>4</sub> cell.</p>
Full article ">
19 pages, 10442 KiB  
Article
Comparison of Bioengineered Scaffolds for the Induction of Osteochondrogenic Differentiation of Human Adipose-Derived Stem Cells
by Elena Fiorelli, Maria Giovanna Scioli, Sonia Terriaca, Arsalan Ul Haq, Gabriele Storti, Marta Madaghiele, Valeria Palumbo, Ermal Pashaj, Fabio De Matteis, Diego Ribuffo, Valerio Cervelli and Augusto Orlandi
Bioengineering 2024, 11(9), 920; https://doi.org/10.3390/bioengineering11090920 - 14 Sep 2024
Viewed by 1145
Abstract
Osteochondral lesions may be due to trauma or congenital conditions. In both cases, therapy is limited because of the difficulty of tissue repair. Tissue engineering is a promising approach that relies on designed scaffolds with variable mechanical attributes to favor cell attachment and [...] Read more.
Osteochondral lesions may be due to trauma or congenital conditions. In both cases, therapy is limited because of the difficulty of tissue repair. Tissue engineering is a promising approach that relies on designed scaffolds with variable mechanical attributes to favor cell attachment and differentiation. Human adipose-derived stem cells (hASCs) are a very promising cell source in regenerative medicine with osteochondrogenic potential. Based on the assumption that stiffness influences cell commitment, we investigated three different scaffolds: a semisynthetic animal-derived GelMA hydrogel, a combined scaffold made of rigid PEGDA coated with a thin GelMA layer and a decellularized plant-based scaffold. We investigated the role of different biomechanical stimulations in the scaffold-induced osteochondral differentiation of hASCs. We demonstrated that all scaffolds support cell viability and spontaneous osteochondral differentiation without any exogenous factors. In particular, we observed mainly osteogenic commitment in higher stiffness microenvironments, as in the plant-based one, whereas in a dense and softer matrix, such as in GelMA hydrogel or GelMA-coated-PEGDA scaffold, chondrogenesis prevailed. We can induce a specific cell commitment by combining hASCs and scaffolds with particular mechanical attributes. However, in vivo studies are needed to fully elucidate the regenerative process and to eventually suggest it as a potential approach for regenerative medicine. Full article
Show Figures

Figure 1

Figure 1
<p>GelMA scaffold characterization and biomechanical test. (<b>A</b>) Hydrogels made with 10% and 15% of GelMA. (<b>B</b>,<b>C</b>) SEM imaging of 10% and 15% GelMA hydrogel, respectively. (<b>D</b>) Stress and strain graph showing a representative elastic modulus for 10% and 15% GelMA hydrogels.</p>
Full article ">Figure 2
<p>PEGDA scaffold production. (<b>A</b>) The final PEGDA scaffold with GelMA coating. (<b>B</b>) SEM images of PEGDA scaffold (GelMA 0.5%) showing its ultrastructure and porosity. (<b>C</b>) Stress and strain graph showing a representative elastic modulus for PEGDA-0,5% GelMA and PEGDA-5% GelMA.</p>
Full article ">Figure 3
<p>(<b>A</b>) Celery-based scaffold after decellularization and sterilization in 70% ethanol. (<b>B</b>) SEM images of decellularized celery-based scaffolds (24 h and 72 h SDS protocol). (<b>C</b>) Representative trends of stress and strain (0–5% strain) comparing different decellularized protocols (24 h and 72 h) and cut orientation (longitudinally cut and transversal cut).</p>
Full article ">Figure 4
<p>Degradation tests. (<b>A</b>) Representative images of 10% and 15% GelMA hydrogel degradation with significant differences at T3 and T4. (<b>B</b>) Graph showing the degradation trend of 10% and 15% GelMA hydrogel in 0.5% collagenase at different points (T, 15 min intervals). (<b>C</b>) Representative images of PEGDA-0.5% GelMA scaffold degradation at T0 (baseline) and T2 (20 h of incubation). (<b>D</b>) Graphs showing the degradation trend of PEGDA scaffold in H2O, 1N NAOH and 1N HCL, and (<b>E</b>) the comparison between PEGDA-0.5% GelMA and PEGDA-5% GelMA in 1N NAOH. The different time points are: T0 (baseline), T1 (18 h of incubation), T2 (20 h of incubation) and T3 (44 h of incubation). (<b>F</b>,<b>G</b>) Degradation test graphs of celery-based scaffolds, (<b>F</b>) cut longitudinally or (<b>G</b>) transversally, incubated with H2O, 1N NAOH and 1N HCL, respectively. T0, T1, T2, T3 and T4 stands for baseline, 2, 4, 6 and 8 weeks. Results are reported as mean ± SEM of <span class="html-italic">n</span> = 3 samples/groups. T test: * <span class="html-italic">p</span> &lt; 0.05 and ** <span class="html-italic">p</span> &lt; 0.001; § <span class="html-italic">p</span> &lt; 0.0001 and §§ <span class="html-italic">p</span> &lt; 0.00001.</p>
Full article ">Figure 5
<p>Swelling tests. (<b>A</b>) Representative images of 10% and 15% GelMA hydrogel before and after the swelling test at T0 (baseline) and T4 (20 min). (<b>B</b>) Graph showing the swelling trend of 10% and 15% GelMA hydrogel, each time point (T) is a 5 min interval. (<b>C</b>) Representative images of PEGDA-0.5% GelMA scaffold before and after the swelling degree at T0 (baseline) and T4 (20 min). (<b>D</b>) Graph showing the swelling trend of PEGDA-0.5% GelMA, PEGDA-5% GelMA and PEGDA-0% GelMA scaffold for each time point (T, 5 min intervals). (<b>E</b>) Representative images of celery-based scaffolds cut longitudinally and transversally dried (T0, baseline) and hydrated (T4, 40 min). (<b>F</b>) Graph showing the swelling trend of celery-based scaffolds cut longitudinally or transversally at different time points (T, 10 min-intervals). Since the standard size of scaffold was too light and did not weigh enough to permit a precise measurement, a bigger scaffold was made purposely to perform this test. Results are reported as mean ± SEM of <span class="html-italic">n</span> = 3 samples/group.</p>
Full article ">Figure 6
<p>hASC survival and differentiation in GelMA hydrogels. (<b>A</b>) CCK8 assay of hASCs encapsulated in the GelMA hydrogel at different time points (1 week intervals). Results are reported as mean ± SEM of <span class="html-italic">n</span> = 3 samples/group. (<b>B</b>) Representative confocal imaging of Dead/live fluorescence staining of hASCs in the GelMA hydrogel after 3 weeks of culture. Red cells are dead while green cells are alive. (<b>C</b>) Alcian blue staining of GelMA hydrogel after 24 h and 3 weeks of culture. (<b>D</b>) Alizarin red staining of GelMA hydrogel after 24 h and 3 weeks of culture. (<b>E</b>,<b>F</b>) Confocal imaging of immunofluorescence for COL2A1 (green) and OCN (red) in hASCs encapsulated in GelMA hydrogel after 3 weeks of culture.</p>
Full article ">Figure 7
<p>hASC survival and differentiation in PEGDA scaffold. (<b>A</b>) CCK8 assay of hASCs seeded in the PEGDA scaffold at different time points (1 week intervals). Results are reported as mean ± SEM of <span class="html-italic">n</span> = 3 samples/group. T test: * <span class="html-italic">p</span> &lt; 0.05. (<b>B</b>) Representative confocal imaging of Dead/live fluorescence staining of hASCs seeded in the PEGDA scaffold after 3 weeks of culture. (<b>C</b>) Confocal imaging of immunofluorescence for COL2A1 (green) and OCN (red) in hASCs seeded in the PEGDA scaffold after 3 weeks of culture.</p>
Full article ">Figure 8
<p>hASC survival and differentiation in celery-based scaffold. (<b>A</b>) SEM imaging of a single hASC inside a niche of the celery-based scaffold. (<b>B</b>) CCK8 assay of hASCs seeded in the celery-based scaffold at different time points (1 week intervals). Results are reported as mean ± SEM of <span class="html-italic">n</span> = 3 samples/group. T test: * <span class="html-italic">p</span> &lt; 0.05. (<b>C</b>) Representative confocal imaging of Dead/live fluorescence staining of hASCs seeded in the scaffold after 3 weeks of culture. (<b>D</b>) Confocal 3D stack and (<b>E</b>) the projection of hASC distribution inside the scaffold. (<b>F</b>) Confocal imaging of immunofluorescence for COL2A1 (green) and OCN (red) in hASCs seeded in the PEGDA scaffold after 3 weeks of culture.</p>
Full article ">
10 pages, 1316 KiB  
Article
A Comparative Study of Ni-Based Catalysts Prepared by Various Sol–Gel Routes
by Atheer Al Khudhair, Karim Bouchmella, Radu Dorin Andrei, Vasile Hulea and Ahmad Mehdi
Molecules 2024, 29(17), 4172; https://doi.org/10.3390/molecules29174172 - 3 Sep 2024
Viewed by 1326
Abstract
The use of heterogeneous catalysts to increase the development of green chemistry is a rapidly growing area of research to save industry money. In this paper, mesoporous SiO2-Al2O3 mixed oxide supports with various Si/Al ratios were prepared using [...] Read more.
The use of heterogeneous catalysts to increase the development of green chemistry is a rapidly growing area of research to save industry money. In this paper, mesoporous SiO2-Al2O3 mixed oxide supports with various Si/Al ratios were prepared using two different sol–gel routes: hydrolytic sol–gel (HSG) and non-hydrolytic sol–gel (NHSG). The HSG route was investigated in both acidic and basic media, while the NHSG was explored in the presence of ethanol and diisopropyl ether as oxygen donors. The resulting SiO2-Al2O3 mixed oxide supports were characterized using EDX, N2 physisorption, powder XRD, 29Si, 27Al MAS-NMR and NH3-TPD. The mesoporous SiO2-Al2O3 supports prepared by NHSG seemed to be more regularly distributed and also more acidic. Consequently, a simple one-step NHSG (ether and alcohol routes) was selected to prepare mesoporous and acidic SiO2-Al2O3-NiO mixed oxide catalysts, which were then evaluated in ethylene oligomerization. The samples prepared by the NHSG ether route showed better activity than those prepared by the NHSG alcohol route in the oligomerization of ethylene at 150 °C. Full article
(This article belongs to the Special Issue Catalysis for Green Chemistry II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p><sup>29</sup>Si CP−MAS NMR spectra of prepared Si<sub>10</sub>Al<sub>90</sub> samples following the HSG in Acidic Media (AM) and basic media (BM) and NHSG using the ether route (ER) and alcohol route (AR) as oxygen donors.</p>
Full article ">Figure 2
<p><sup>27</sup>Al MAS NMR spectra of prepared Si<sub>0</sub>Al<sub>100</sub> samples following HSG in Acidic Media (AM) and basic media (BM) and NHSG using ether route (ER) and alcohol route (AR) as oxygen donors.</p>
Full article ">Figure 3
<p><sup>27</sup>Al MAS NMR spectra of prepared Si<sub>10</sub>Al<sub>90</sub> samples following HSG in Acidic Media (AM) and basic media (BM) and NHSG using ether route (ER) and alcohol route (AR) as oxygen donors.</p>
Full article ">Scheme 1
<p>Experimental conditions for materials syntheses following the hydrolytic sol-gel process (HSG) in acidic Media (AM) and basic media (BM) and non-hydrolytic sol-gel using the ether route (ER) and alcohol route (AR) as oxygen donors.</p>
Full article ">Scheme 2
<p>Reaction equation of the formation of C4, C6 and C8 olefins.</p>
Full article ">
17 pages, 7180 KiB  
Article
Development of Cerium Oxide-Laden GelMA/PCL Scaffolds for Periodontal Tissue Engineering
by Sahar Aminmansour, Lais M. Cardoso, Caroline Anselmi, Ana Beatriz Gomes de Carvalho, Maedeh Rahimnejad and Marco C. Bottino
Materials 2024, 17(16), 3904; https://doi.org/10.3390/ma17163904 - 7 Aug 2024
Viewed by 1068
Abstract
This study investigated gelatin methacryloyl (GelMA) and polycaprolactone (PCL) blend scaffolds incorporating cerium oxide (CeO) nanoparticles at concentrations of 0%, 5%, and 10% w/w via electrospinning for periodontal tissue engineering. The impact of photocrosslinking on these scaffolds was evaluated by comparing crosslinked (C) [...] Read more.
This study investigated gelatin methacryloyl (GelMA) and polycaprolactone (PCL) blend scaffolds incorporating cerium oxide (CeO) nanoparticles at concentrations of 0%, 5%, and 10% w/w via electrospinning for periodontal tissue engineering. The impact of photocrosslinking on these scaffolds was evaluated by comparing crosslinked (C) and non-crosslinked (NC) versions. Methods included Fourier transform infrared spectroscopy (FTIR) for chemical analysis, scanning electron microscopy (SEM) for fiber morphology/diameters, and assessments of swelling capacity, degradation profile, and biomechanical properties. Biological evaluations with alveolar bone-derived mesenchymal stem cells (aBMSCs) and human gingival fibroblasts (HGFs) encompassed tests for cell viability, mineralized nodule deposition (MND), and collagen production (CP). Statistical analysis was performed using Kruskal–Wallis or ANOVA/post-hoc tests (α = 5%). Results indicate that C scaffolds had larger fiber diameters (~250 nm) compared with NC scaffolds (~150 nm). NC scaffolds exhibited higher swelling capacities than C scaffolds, while both types demonstrated significant mass loss (~50%) after 60 days (p < 0.05). C scaffolds containing CeO showed increased Young’s modulus and tensile strength than NC scaffolds. Cells cultured on C scaffolds with 10% CeO exhibited significantly higher metabolic activity (>400%, p < 0.05) after 7 days among all groups. Furthermore, CeO-containing scaffolds promoted enhanced MND by aBMSCs (>120%, p < 0.05) and increased CP in 5% CeO scaffolds for both variants (>180%, p < 0.05). These findings underscore the promising biomechanical properties, biodegradability, cytocompatibility, and enhanced tissue regenerative potential of CeO-loaded GelMA/PCL scaffolds for periodontal applications. Full article
(This article belongs to the Special Issue Advanced Materials for Oral Application (3rd Edition))
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the electrospinning technique used for the fabrication of the designed scaffolds.</p>
Full article ">Figure 2
<p>Fourier transform infrared (FTIR) spectra of the non-crosslinked (<b>a</b>) and crosslinked (<b>b</b>) GelMA/PCL nanofibrous scaffolds containing different concentrations of cerium oxide (CeO) (0–control; 5, and 10% <span class="html-italic">w/w</span>). FTIR spectra of the formulated scaffolds immersed in PBS after 3, 7, and 10 days (<b>c</b>).</p>
Full article ">Figure 3
<p>SEM images captured from the surfaces of non-crosslinked and crosslinked scaffolds at a magnification of 7000×. Histograms represent the frequency distribution (%) of fiber diameters studied from 150 fibers. Boxplots show the median fiber diameter values (25th–75th percentiles). Groups identified by different letters show a statistically significant difference (Kruskal–Wallis, followed by Dunn’s post-hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 4
<p>Degradation profile (<b>a</b>,<b>c</b>) (remaining mass; n = 8) in PBS containing 1 U/mL collagenase and swelling capacity (<b>b</b>,<b>d</b>) (n = 5) in PBS at 37 °C of non-crosslinked (NC) (<b>a</b>,<b>b</b>) and crosslinked (C) (<b>c</b>,<b>d</b>) GelMA/PCL scaffolds containing CeO (0—control; 5, and 10% <span class="html-italic">w/w</span>) at different timepoints. Geometric symbols are means, and error bars represent 95% confidence intervals (α = 5%).</p>
Full article ">Figure 5
<p>Mechanical characterizations of the designed scaffolds. (<b>a</b>,<b>b</b>) Stress–strain diagram, (<b>c</b>) Young’s modulus (MPa), (<b>d</b>) tensile strength (MPa), and (<b>e</b>) elongation at break (%) under non-crosslinked (NC) and crosslinked (C) conditions. Columns represent mean values, and error bars represent standard deviations. Statistically significant differences between groups within each condition are denoted by different capital letters, while significant differences between conditions within each group are indicated by different lowercase letters. Analysis was conducted using two-way ANOVA, followed by Sidak’s post-hoc test, with a significance level set at <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 6
<p>Viability (% of control—NC GelMA/PCL) of alveolar bone-derived mesenchymal stem cells (<b>a</b>) and human gingival fibroblasts (<b>b</b>) seeded on scaffold surfaces after 1, 3, and 7 days (n = 8) was assessed. Columns represent mean values, and error bars denote standard deviations. Significant differences between groups within each time point are indicated by different capital letters, while significant differences between time points within each group are denoted by different lowercase letters (two-way repeated measures ANOVA, followed by Sidak’s post-hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 7
<p>Collagen synthesis (% of control—NC GelMA/PCL) (<b>a</b>) by human gingival fibroblasts seeded on the surface of the scaffolds after culturing for 3 days in serum-free media (Sircol assay) (n = 6). Columns depict mean values, with error bars representing standard deviations. Groups identified by different letters are statistically different from each other (Welch’s ANOVA, followed by Games–Howell’s post-hoc test, <span class="html-italic">p</span> &lt; 0.05). Quantitative (<b>b</b>) (% of control—NC GelMA/PCL) and qualitative (<b>c</b>) (scale bar 1000 µm) analysis of mineral nodule deposition (Alizarin Red assay) by alveolar bone-derived mesenchymal stem cells seeded on scaffold surfaces after 14 and 21 days of cell culture (n = 8). Columns represent mean values, and error bars denote standard deviations. Significant differences between groups within each time point are indicated by different capital letters, while significant differences between time points within each group are denoted by different lowercase letters (Two-way ANOVA, followed by Sidak’s post-hoc test, <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
13 pages, 1924 KiB  
Article
Synthesis of Gelatin Methacryloyl Analogs and Their Use in the Fabrication of pH-Responsive Microspheres
by Karolina Valente, Geneviève N. Boice, Cameron Polglase, Roman G. Belli, Elaina Bourque, Afzal Suleman and Alexandre Brolo
Pharmaceutics 2024, 16(8), 1016; https://doi.org/10.3390/pharmaceutics16081016 - 31 Jul 2024
Cited by 1 | Viewed by 1229
Abstract
pH-responsive hydrogels have numerous applications in tissue engineering, drug delivery systems, and diagnostics. Gelatin methacryloyl (GelMA) is a biocompatible, semi-synthetic polymer prepared from gelatin. When combined with aqueous solvents, GelMA forms hydrogels that have extensive applications in biomedical engineering. GelMA can be produced [...] Read more.
pH-responsive hydrogels have numerous applications in tissue engineering, drug delivery systems, and diagnostics. Gelatin methacryloyl (GelMA) is a biocompatible, semi-synthetic polymer prepared from gelatin. When combined with aqueous solvents, GelMA forms hydrogels that have extensive applications in biomedical engineering. GelMA can be produced with different degrees of methacryloyl substitution; however, the synthesis of this polymer has not been tuned towards producing selectively modified materials for single-component pH-responsive hydrogels. In this work, we have explored two different synthetic routes targeting different gelatin functional groups (amine, hydroxyl, and/or carboxyl) to produce two GelMA analogs: gelatin A methacryloyl glycerylester (polymer A) and gelatin B methacrylamide (polymer B). Polymers A and B were used to fabricate pH-responsive hydrogel microspheres in a flow-focusing microfluidic device. At neutral pH, polymer A and B microspheres displayed an average diameter of ~40 µm. At pH 6, microspheres from polymer A showed a swelling ratio of 159.1 ± 11.5%, while at pH 10, a 288.6 ± 11.6% swelling ratio was recorded for polymer B particles. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Synthesis of polymers A and B. (<b>A</b>) Polymer A was produced by reacting gelatin A with glycidyl methacrylate at pH 3.5. After a total reaction time of 24 h, 100 mL of pH 3.5 HCl solution was added to the aqueous mixture. The final solution was then dialyzed, followed by lyophilization to obtain polymer A. (<b>B</b>) Polymer B was produced by reacting gelatin B with methacrylic anhydride. After a total reaction time of 3 h, 100 mL of PBS was added to the solution, followed by dialysis and lyophilization.</p>
Full article ">Figure 2
<p>Synthesis of polymers A and B. (<b>A</b>) Two synthesis routes from gelatin. The reaction of gelatin A with glycidyl methacrylate at pH 3.5 produced polymer A (left), while the reaction of gelatin B with methacrylic anhydride at pH 7.4 generated polymer B (right). (<b>B</b>) Mechanism of methacrylation of gelatin type A hydroxyl groups with glycidyl methacrylate through ring-opening reaction. (<b>C</b>) <sup>1</sup>H NMR spectra of polymer A (top) and gelatin type A (bottom). (<b>D</b>) Mechanism of methacrylation of gelatin type B with methacrylic anhydride. (<b>E</b>) <sup>1</sup>H NMR spectra of polymer B (top) and gelatin type B (bottom).</p>
Full article ">Figure 3
<p>Fabrication of GelMA microspheres. (<b>A</b>) Microspheres were fabricated inside of a flow-focusing microfluidic device. (<b>B</b>) The device contained two phases (continuous and dispersed). The dispersed phase was composed of 5% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) GelMA (polymer A or B), while the continuous phase consisted of mineral oil with Span 80. (<b>C</b>) Flow profile during production of microspheres. (<b>D</b>) Optical microscopy (OM) image of polymer A microspheres. (<b>E</b>) OM image of polymer B microspheres.</p>
Full article ">Figure 4
<p>Swelling behavior of GelMA microspheres. (<b>A</b>) Polymer A microspheres displayed swelling at acidic pH (6.0), while shrinkage was seen at basic pH (10). (<b>B</b>) Swelling and shrinking of polymer A microspheres was also investigated by adding red-fluorescent PS nanoparticles to the hydrogel solution. The same swelling and de-swelling behaviors were observed for the fluorescent microspheres. (<b>C</b>) The opposite behavior was observed for polymer B microspheres, in which swelling was observed at basic pH, while shrinkage was seen in an acidic environment. (<b>D</b>) Microspheres were also fabricated by adding PS particles to polymer B pre-hydrogel solution, providing easier observation of the increase and decrease in the diameter of the microspheres.</p>
Full article ">Figure 5
<p>Swelling and de-swelling behavior of GelMA microspheres. (<b>A</b>) Change in diameter was observed for polymers A and B, according to the pH of the environment. For polymer A, a decrease in diameter was seen with an increase in pH, while the opposite behavior was observed for polymer B (increase in diameter with increase in pH). (<b>B</b>) Swelling ratios of 159.1 ± 11.5% and 288.6 ± 11.6% were obtained for polymers A and B, respectively, at pH 6 and pH 10. Shrinking ratios of 62.3 ± 7.6% and 58.4 ± 1.2% were observed for polymer A at pH 10 and for polymer B at pH 6, respectively.</p>
Full article ">
Back to TopTop