[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

Search Results (57)

Search Parameters:
Keywords = 8-arm-polyethylene glycol

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 3866 KiB  
Article
Preparation and Rheological Evaluation of Thiol–Maleimide/Thiol–Thiol Double Self-Crosslinking Hyaluronic Acid-Based Hydrogels as Dermal Fillers for Aesthetic Medicine
by Chia-Wei Chu, Wei-Jie Cheng, Bang-Yu Wen, Yu-Kai Liang, Ming-Thau Sheu, Ling-Chun Chen and Hong-Liang Lin
Gels 2024, 10(12), 776; https://doi.org/10.3390/gels10120776 - 28 Nov 2024
Viewed by 490
Abstract
This study presents the development of thiol–maleimide/thiol–thiol double self-crosslinking hyaluronic acid-based (dscHA) hydrogels for use as dermal fillers. Hyaluronic acid with varying degrees of maleimide substitution (10%, 20%, and 30%) was synthesized and characterized, and dscHA hydrogels were fabricated using [...] Read more.
This study presents the development of thiol–maleimide/thiol–thiol double self-crosslinking hyaluronic acid-based (dscHA) hydrogels for use as dermal fillers. Hyaluronic acid with varying degrees of maleimide substitution (10%, 20%, and 30%) was synthesized and characterized, and dscHA hydrogels were fabricated using two molecular weights of four-arm polyethylene glycol (PEG10K/20K)–thiol as crosslinkers. The six resulting dscHA hydrogels demonstrated solid-like behavior with distinct physical and rheological properties. SEM analysis revealed a decrease in porosity with higher crosslinker MW and maleimide substitution. The swelling ratios of the six hydrogels reached equilibrium at approximately 1 h and ranged from 20% to 35%, indicating relatively low swelling. Degradation rates decreased with increasing maleimide substitution, while crosslinker MW had little effect. Higher maleimide substitution also required greater injection force. Elastic modulus (G′) in the linear viscoelastic region increased with maleimide substitution and crosslinker MW, indicating enhanced firmness. All hydrogels displayed similar creep-recovery behavior, showing instantaneous deformation under constant stress. Alternate-step strain tests indicated that all six dscHA hydrogels could maintain elasticity, allowing them to integrate with the surrounding tissue via viscous deformation caused by the stress exerted by changes in facial expression. Ultimately, the connection between the clinical performance of the obtained dscHA hydrogels used as dermal filler and their physicochemical and rheological properties was discussed to aid clinicians in the selection of the most appropriate hydrogel for facial rejuvenation. While these findings are promising, further studies are required to assess irritation, toxicity, and in vivo degradation before clinical use. Overall, it was concluded that all six dscHA hydrogels show promise as dermal fillers for various facial regions. Full article
(This article belongs to the Special Issue Recent Research on Medical Hydrogels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>1H NMR (<b>A</b>) and FTIR spectra (<b>B</b>) of HA and HA-Mal with three different degrees of substitution of maleimide on HA (HM10, HM20, and HM30).</p>
Full article ">Figure 2
<p>Reaction scheme illustrating the formation of <span class="html-italic">dsc</span>HA hydrogels.</p>
Full article ">Figure 3
<p>SEM images of six <span class="html-italic">dsc</span>HA hydrogels (HM10-4SH10K (<b>A</b>), HM20-4SH10K (<b>B</b>), HM30-4SH10K (<b>C</b>), HM10-4SH20K (<b>D</b>), HM20-4SH20K (<b>E</b>), and HM30-4SH20K (<b>F</b>). (Scale bar: 500 µm).</p>
Full article ">Figure 4
<p>The swelling ratio profiles (<b>A</b>) and degradation profiles (<b>B</b>) for HAs with various levels of maleimide substitution and thiol-containing crosslinkers with two different MWs (designated as HM10-4SH10K, HM10-4SH20K, HM20-4SH10K, HM20-4SH20K, HM30-4SH10K, and HM30-4SH20K, respectively).</p>
Full article ">Figure 5
<p>Injection force through a 26 G needle measured for six <span class="html-italic">dsc</span>HA hydrogels (HM10-4SH10K, HM10-4SH20K, HM20-4SH10K, HM20-4SH20K, HM30-4SH10K, and HM30-4SH20K).</p>
Full article ">Figure 6
<p>Rheological evaluation of <span class="html-italic">dsc</span>HA hydrogels. Amplitude sweep (<b>A</b>) and frequency sweep (<b>B</b>) of the six <span class="html-italic">dsc</span>HA hydrogels, showing the linear viscoelastic (LVE) region and gel behavior. Tan δ values (<b>C</b>) of the six <span class="html-italic">dsc</span>HA hydrogels, indicating whether the behavior is elastic-dominant or viscous-dominant.</p>
Full article ">Figure 7
<p>Creep-recovery experiments (constant stress) were performed with an applied shear stress of 5 Pa for 10 min followed by 20 min of recovery (<b>A</b>), and alternate-step strain tests with five repetitions of shear-stress application and relaxation (<b>B</b>) were performed to study the deformation and recovery of the hydrogel network.</p>
Full article ">
15 pages, 5902 KiB  
Article
In Situ Crosslinked Biodegradable Hydrogels Based on Poly(Ethylene Glycol) and Poly(ε-Lysine) for Medical Application
by Xia Ding, Bing Yang and Zhaosheng Hou
Molecules 2024, 29(22), 5435; https://doi.org/10.3390/molecules29225435 - 18 Nov 2024
Viewed by 663
Abstract
Hydrogels have emerged as promising biomaterials due to their excellent performance; however, their biocompatibility, biodegradability, and absorbability still require improvement to support a broader range of medical applications. This paper presents a new biofunctionalized hydrogel based on in situ crosslinking between maleimide-terminated four-arm-poly(ethylene [...] Read more.
Hydrogels have emerged as promising biomaterials due to their excellent performance; however, their biocompatibility, biodegradability, and absorbability still require improvement to support a broader range of medical applications. This paper presents a new biofunctionalized hydrogel based on in situ crosslinking between maleimide-terminated four-arm-poly(ethylene glycol) (4–arm–PEG–Mal) and poly(ε-lysine) (ε–PL). The PEG/ε–PL hydrogels, named LG–n, were rapidly formed via amine/maleimide reaction by mixing 4–arm–PEG–Mal and ε–PL under physiological conditions. The corresponding dry gels (DLG–n) were obtained through a freeze-drying technique. 1H NMR, FT–IR, and SEM were utilized to confirm the structures of 4–arm–PEG–Mal and LG–n (or DLG–n), and the effects of solid content on the physicochemical properties of the hydrogels were investigated. Although high solid content could increase the swelling ratio, all LG–n samples exhibited a low equilibrium swelling ratio of less than 30%. LG–7, which contained moderate solid content, exhibited optimal compression properties characterized by a compressive fracture strength of 45.2 kPa and a deformation of 69.5%. Compression cycle tests revealed that LG–n demonstrated good anti-fatigue performance. In vitro degradation studies confirmed the biodegradability of LG–n, with the degradation rate primarily governing the drug (ceftibuten) release efficiency, leading to a sustained release duration of four weeks. Cytotoxicity tests, cell survival morphology observation, live/dead assays, and hemolysis tests indicated that LG–n exhibited excellent cytocompatibility and low hemolysis rates (<5%). Furthermore, the broad-spectrum antibacterial activity of LG–n was verified by an inhibition zone method. In conclusion, the developed LG–n hydrogels hold promising applications in the medical field, particularly as drug sustained-release carriers and wound dressings. Full article
(This article belongs to the Special Issue Hydrogels: Preparation, Characterization, and Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Photo of the dual–chamber syringe equipped with a helical mixing head.</p>
Full article ">Figure 2
<p>FT–IR spectra for (<b>a</b>) ε–PL, (<b>b</b>) 4–arm–PEG–Mal, and (<b>c</b>) DLG–7.</p>
Full article ">Figure 3
<p>SEM photos of (<b>a</b>) DLG–3, (<b>b</b>) DLG–7, (<b>c</b>) DLG–11 (scale bar: 500 μm).</p>
Full article ">Figure 4
<p>(<b>a</b>) TGA and (<b>b</b>) DTG curves of ε–PL, 4–arm–PEG–Mal, and DLG–n.</p>
Full article ">Figure 5
<p>DSC curves of ε–PL, 4–arm–PEG–Mal, and DLG–n.</p>
Full article ">Figure 6
<p>Swelling curves of ε–PL, 4–arm–PEG–Mal, and DLG–n (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 7
<p>(<b>a</b>) Compressive stress–strain profiles of LG–n hydrogels and (<b>b</b>) cyclic compressive stress–strain profiles of LG–7 hydrogels.</p>
Full article ">Figure 8
<p>Mass loss curves of LG–n against degradation time in PBS (pH 7.4) at 37 °C (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 9
<p>Drug release curves from LG–n hydrogels in PBS (pH 7.4) at 37 °C (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 10
<p>Cell survival rate in LG–n hydrogel extracts by MTT (37 °C, 72 h, <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 11
<p>(<b>a</b>) Cell morphologies on LG–n hydrogel surface (37 °C, 72 h, scale bar: 100 μm) and (<b>b</b>) live/dead cells in LG–n extracts (37 °C, 72 h, scale bar: 100 μm).</p>
Full article ">Figure 12
<p>Hemolysis test images and hemolysis rates of hydrogel LG–n (37 °C, 24 h, <span class="html-italic">n</span> = 3).</p>
Full article ">Figure 13
<p>Antibacterial activities of LG–n against <span class="html-italic">E. coli</span> and <span class="html-italic">S. aureus</span>.</p>
Full article ">Figure 14
<p>(<b>a</b>) Synthesis of 4–arm–PEG–Mal, (<b>b</b>) preparation of hydrogel LG–n, and (<b>c</b>) crosslinking schematic diagram of hydrogel LG–n.</p>
Full article ">
18 pages, 2390 KiB  
Article
Paclitaxel-Loaded, Pegylated Carboxylic Graphene Oxide with High Colloidal Stability, Sustained, pH-Responsive Release and Strong Anticancer Effects on Lung Cancer A549 Cell Line
by Athina Angelopoulou, Myria Papachristodoulou, Efstathia Voulgari, Andreas Mouikis, Panagiota Zygouri, Dimitrios P. Gournis and Konstantinos Avgoustakis
Pharmaceutics 2024, 16(11), 1452; https://doi.org/10.3390/pharmaceutics16111452 - 14 Nov 2024
Viewed by 1006
Abstract
Background: Graphene Oxide (GO) has shown great potential in biomedical applications for cancer therapeutics. The biosafety and stability issues of GO in biological media have been addressed by functionalization with polyethylene glycol (PEG). Methods: In this work, carboxylated, nanosized GO (nCGO) [...] Read more.
Background: Graphene Oxide (GO) has shown great potential in biomedical applications for cancer therapeutics. The biosafety and stability issues of GO in biological media have been addressed by functionalization with polyethylene glycol (PEG). Methods: In this work, carboxylated, nanosized GO (nCGO) was evaluated as a potential carrier of paclitaxel (PCT). The effect of PEG characteristics on particle size and surface charge, colloidal stability, drug, and release, and the hemolytic potential of nCGO, was investigated. Optimum PEG-nCGO/PCT formulations based on the above properties were evaluated for their anticancer activity (cytotoxicity and apoptosis induction) in the A549 lung cancer cell line. Results: An increase in the length of linear PEG chains and the use of branched (4-arm) instead of linear PEG resulted in a decrease in hydrodynamic diameter and an increase in ζ potential of the pegylated nCGO particles. Pegylated nCGO exhibited high colloidal stability in phosphate-buffered saline and in cell culture media and low hemolytic effect, even at a relatively high concentration of 1 mg/mL. The molecular weight of PEG and branching adversely affected PCT loading. An increased rate of PCT release at an acidic pH of 6.0 compared to the physiological pH of 7.4 was observed with all types of pegylated nCGO/PCT. Pegylated nCGO exhibited lower cytotoxicity and apoptotic activity than non-pegylated nCGO. Cellular uptake of pegylated nCGO increased with incubation time with cells leading to increased cytotoxicity of PEG-nCGO/PCT with incubation time, which became higher than that of free PCT at 24 and 48 h of incubation. Conclusions: The increased biocompatibility of the pegylated nCGO and the enhanced anticancer activity of PEG-nCGO/PCT compared to free PCT are desirable properties with regard to the potential clinical application of PEG-nCGO/PCT as an anticancer nanomedicine. Full article
Show Figures

Figure 1

Figure 1
<p>Characterization of nCGO-PEG particles: (<b>A</b>) FTIR spectra of nCGO (black) mPEG-NH<sub>2</sub> (amine) polymer (light blue), and nCGO-PEG particles of varied MW (2, 10, 20 kDa) (red, blue magenta); (<b>B</b>) thermograms of nCGO, mPEG(10 kDa)-NH<sub>2</sub>, and nCGO-PEG(10 kDa) up to 600 °C; (<b>C</b>) SEM micrographs of nCGO-PEG(10 kDa) at scale bar of 200 nm (<b>C</b>), 50 nm (<b>D</b>), and 100 nm (<b>E</b>).</p>
Full article ">Figure 2
<p>Colloidal stability of the nCGO and nCGO-PEG particles exhibiting (<b>A</b>) average size distribution by DLS and (<b>B</b>) distribution of ζ-potential for a period of 4 weeks. The stability of nCGO and nCGO-PEG particles in RPMI and PBS media as presented by (<b>C</b>) average size and (<b>D</b>) ζ-potential at 5, 54, and 48 h. The statistical significance is ** <span class="html-italic">p</span> &lt; 0.001, *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 3
<p>(<b>A</b>) Hemolysis observed at varied concentrations of the nCGO and nCGO-PEG particles. (<b>B</b>) Representative hemolysis photographs at a particle concentration of 25 μg/mL and with the control (positive, negative) samples. The statistical significance is *** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 4
<p>Paclitaxel release profile from the PCT/nCGO-PEG particles in PBS buffer with pH (<b>A</b>) 7.4 and (<b>B</b>) 6.0.</p>
Full article ">Figure 5
<p>Evaluation of the nCGO-PEG(10 kDa) particles against lung adenocarcinoma A549 cell line for the induced anticancer effect (cytotoxicity) at (<b>A</b>) 24 h and (<b>B</b>) 48 h. Internalization of FITC-labeled nCGO-PEG(10 kDa) in comparison with nCGO-particles (<b>C</b>) and Fluorescence microscopy by PI post-fixation staining method of A549 cellular nuclei (<b>D</b>) and cells treated with FITC-labeled nCGO-PEG(10 kDa) particles (<b>E</b>). Evaluation on programmed cell death of A549 cells by apoptosis assay induced by nCGO (grey circles), nCGO-PEG(10 kDa) blank (blue circles), PCT (orange circle), and nCGO-PEG(10 kDa)/PCT loaded (yellow circle) particles (<b>F</b>). The statistical significance is * <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 ">
6 pages, 2005 KiB  
Proceeding Paper
Fibroblast and THP-1 Cell Response to Multi-Arm PEGNHS-Modified Decellularized Porcine Pericardium
by Sreypich Say, Mika Suzuki, Yoshihide Hashimoto, Tsuyoshi Kimura and Akio Kishida
Mater. Proc. 2024, 19(1), 3; https://doi.org/10.3390/materproc2024019003 - 1 Nov 2024
Viewed by 650
Abstract
The adhesion between an implant and a wound could result in over-bleeding when attempting to separate the two. To address this issue, a cell-repelling implant is preferred. In this study, a cell-repelling membrane was prepared by modifying decellularized porcine pericardium with multi-arm polyethylene [...] Read more.
The adhesion between an implant and a wound could result in over-bleeding when attempting to separate the two. To address this issue, a cell-repelling implant is preferred. In this study, a cell-repelling membrane was prepared by modifying decellularized porcine pericardium with multi-arm polyethylene glycol. With this modification technology, we switched the surface properties of the decellularized porcine pericardium from cell-adhering to cell-repelling. The result showed that this pericardium was successfully modified without any effect on the original properties of the pericardium and also maintained a low inflammatory response. The level of cell adhesion on the surface of the membrane was significantly reduced. Full article
(This article belongs to the Proceedings of The 1st International Online Conference on Functional Biomaterials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) H and E staining; (<b>b</b>) DNA residue of porcine pericardium and decellularized porcine pericardium. Data are expressed as the mean ± S.D. ** <span class="html-italic">p</span> &lt; 0.01, where the values for the modified porcine pericardium samples are compared with dP, respectively. The numbers 2,4,8 are the PEG arms, dP: decellularized porcine pericardium, 1,2: NHS ratio. Scale: 100 μm.</p>
Full article ">Figure 2
<p>(<b>a</b>) The amine index; (<b>b</b>) ATR-FITR spectra (<b>c</b>) Immunohistochemistry staining of the non-modification and modification decellularized porcine pericardium. Data are expressed as the mean ± S.D. ** <span class="html-italic">p</span> &lt; 0.01, where the values for the modified porcine pericardium samples are compared with dP, respectively. The numbers 2,4,8 are the PEG arms, dP: decellularized porcine pericardium, 1,2: NHS ratio. Scale: 200 μm.</p>
Full article ">Figure 3
<p>(<b>a</b>) The inflammatory response of the macrophage differentiated from the THP-1 tagged HiBiT; (<b>b</b>) Fibroblast density adherence to the non-modification and modification dP samples, <span class="html-italic">n</span> = 3. Data are expressed as the mean ± S.D.** <span class="html-italic">p</span> &lt; 0.01, where the values for the modified porcine pericardium samples are compared with the dP, respectively. The numbers 2,4,8 are the PEG arms, dP: decellularized porcine pericardium, 1,2: NHS ratio. a,b,c,d,e denote statistically significant differences.</p>
Full article ">
18 pages, 5230 KiB  
Article
Crosslinked Biodegradable Hybrid Hydrogels Based on Poly(ethylene glycol) and Gelatin for Drug Controlled Release
by Zhenzhen Zhao, Zihao Qin, Tianqing Zhao, Yuanyuan Li, Zhaosheng Hou, Hui Hu, Xiaofang Su and Yanan Gao
Molecules 2024, 29(20), 4952; https://doi.org/10.3390/molecules29204952 - 19 Oct 2024
Cited by 1 | Viewed by 1106
Abstract
A series of hybrid hydrogels of poly(ethylene glycol) (PEG) were synthesized using gelatin as a crosslinker and investigated for controlled delivery of the first-generation cephalosporin antibiotic, Cefazedone sodium (CFD). A commercially available 4-arm-PEG–OH was first modified to obtain four-arm-PEG–succinimidyl glutarate (4-arm-PEG–SG), which formed [...] Read more.
A series of hybrid hydrogels of poly(ethylene glycol) (PEG) were synthesized using gelatin as a crosslinker and investigated for controlled delivery of the first-generation cephalosporin antibiotic, Cefazedone sodium (CFD). A commercially available 4-arm-PEG–OH was first modified to obtain four-arm-PEG–succinimidyl glutarate (4-arm-PEG–SG), which formed the gelatin–PEG composite hydrogels (SnNm) through crosslinking with gelatin. To regulate the drug delivery, SnNm hydrogels with various solid contents and crosslinking degrees were prepared. The effect of solid contents and crosslinking degrees on the thermal, mechanical, swelling, degradation, and drug release properties of the hydrogels were intensively investigated. The results revealed that increasing the crosslinking degree and solid content of SnNm could not only enhance the thermal stability, swelling ratio (SR), and compression resistance capacity of SnNm but also prolong the degradation and drug release times. The release kinetics of the SnNm hydrogels were found to follow the first-order model, suggesting that the transport rate of CFD within the matrix of hydrogels is proportional to the concentration of the drug where it is located. Specifically, S1N1-III showed 90% mass loss after 60 h of degradation and a sustained release duration of 72 h. The cytotoxicity assay using the MTT method revealed that cell viability rates of S1N1 were higher than 95%, indicating excellent cytocompatibility. This study offers new insights and methodologies for the development of hydrogels as biomedical composite materials. Full article
(This article belongs to the Special Issue Recent Advances in Porous Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Synthesis of 4-arm-PEG–SG; (<b>b</b>) preparation of S<sub>n</sub>N<sub>m</sub> hydrogel; (<b>c</b>) possible crosslinking mechanism of hydrogel; and (<b>d</b>) optical image of hydrogel.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra for (<b>a</b>) 4-arm-PEG–OH, (<b>b</b>) 4-arm-PEG–GA, and (<b>c</b>) 4-arm-PEG–SG with CDCl<sub>3</sub> as solvent.</p>
Full article ">Figure 3
<p>FT–IR spectra of (<b>a</b>) 4-arm-PEG–SG, (<b>b</b>) DS<sub>2</sub>N<sub>1</sub>–I, (<b>c</b>) DS<sub>1</sub>N<sub>1</sub>–I, (<b>d</b>) DS<sub>1</sub>N<sub>2</sub>–I, and (<b>e</b>) gelatin.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>) DS<sub>2</sub>N<sub>1</sub>–I, (<b>b</b>) DS<sub>1</sub>N<sub>2</sub>–I, and (<b>c</b>) DS<sub>1</sub>N<sub>1</sub>–I. (scale bar: 200 μm).</p>
Full article ">Figure 5
<p>TGA curves of 4-arm-PEG–SG, DS<sub>n</sub>N<sub>n</sub>–I, and gelatin.</p>
Full article ">Figure 6
<p>DSC curves of 4-arm-PEG–SG, DS<sub>n</sub>N<sub>n</sub>–I, and gelatin.</p>
Full article ">Figure 7
<p>Swelling ratio of S<sub>n</sub>N<sub>m</sub> hydrogels in water at 37 °C.</p>
Full article ">Figure 8
<p>(<b>a</b>) Compressive stress–strain curve of S<sub>n</sub>N<sub>m</sub> hydrogels and (<b>b</b>) cyclic compressive stress–strain curves of S<sub>1</sub>N<sub>1</sub>–II.</p>
Full article ">Figure 9
<p>Mass loss curves of S<sub>n</sub>N<sub>m</sub> hydrogels against the degradation time in PBS (pH 7.4) at 37 °C.</p>
Full article ">Figure 10
<p>(<b>a</b>) Drug release profiles from S<sub>n</sub>N<sub>m</sub> hydrogels in PBS (pH 7.4) at 37 °C; release kinetics of (<b>b</b>) S<sub>1</sub>N<sub>1</sub>-III, (<b>c</b>) S<sub>1</sub>N<sub>1</sub>-II, (<b>d</b>) S<sub>1</sub>N<sub>1</sub>-I, (<b>e</b>) S<sub>2</sub>N<sub>1</sub>-I, and (<b>f</b>) S<sub>1</sub>N<sub>2</sub>-I.</p>
Full article ">Figure 11
<p>Cell survival rate in S<sub>n</sub>N<sub>m</sub> hydrogel extracts using MTT at 37 °C for 72 h.</p>
Full article ">Figure 12
<p>(<b>a</b>) Cell morphologies on S<sub>1</sub>N<sub>1</sub>–I hydrogel surface and (<b>b</b>) live/dead cells in S<sub>1</sub>N<sub>1</sub>–I extracts (37 °C, 72 h).</p>
Full article ">
13 pages, 10019 KiB  
Protocol
A Scalable Method to Fabricate 2D Hydrogel Substrates for Mechanobiology Studies with Independent Tuning of Adhesiveness and Stiffness
by Alessandro Gandin, Veronica Torresan, Tito Panciera and Giovanna Brusatin
Methods Protoc. 2024, 7(5), 75; https://doi.org/10.3390/mps7050075 - 26 Sep 2024
Viewed by 856
Abstract
Mechanical signals from the extracellular matrix are crucial in guiding cellular behavior. Two-dimensional hydrogel substrates for cell cultures serve as exceptional tools for mechanobiology studies because they mimic the biomechanical and adhesive characteristics of natural environments. However, the interdisciplinary knowledge required to synthetize [...] Read more.
Mechanical signals from the extracellular matrix are crucial in guiding cellular behavior. Two-dimensional hydrogel substrates for cell cultures serve as exceptional tools for mechanobiology studies because they mimic the biomechanical and adhesive characteristics of natural environments. However, the interdisciplinary knowledge required to synthetize and manipulate these biomaterials typically restricts their widespread use in biological laboratories, which may not have the material science expertise or specialized instrumentation. To address this, we propose a scalable method that requires minimal setup to produce 2D hydrogel substrates with independent modulation of the rigidity and adhesiveness within the range typical of natural tissues. In this method, norbornene-terminated 8-arm polyethylene glycol is stoichiometrically functionalized with RGD peptides and crosslinked with a di-cysteine terminated peptide via a thiol–ene click reaction. Since the synthesis process significantly influences the final properties of the hydrogels, we provide a detailed description of the chemical procedure to ensure reproducibility and high throughput results. We demonstrate examples of cell mechanosignaling by monitoring the activation state of the mechanoeffector proteins YAP/TAZ. This method effectively dissects the influence of biophysical and adhesive cues on cell behavior. We believe that our procedure will be easily adopted by other cell biology laboratories, improving its accessibility and practical application. Full article
(This article belongs to the Section Molecular and Cellular Biology)
Show Figures

Figure 1

Figure 1
<p>Schematic protocol workflow for hydrogel synthesis. Taken from [<a href="#B22-mps-07-00075" class="html-bibr">22</a>] with minor modification.</p>
Full article ">Figure 2
<p>(<b>a</b>) hydrophilic glass slide; (<b>b</b>) wash with NaOH 3 M; (<b>c</b>) dehydration of the substrate; (<b>d</b>) activation of the surface with plasma cleaner; (<b>e</b>) substrate silanization using Repel-silane; (<b>f</b>) hydrophobic glass substrate.</p>
Full article ">Figure 3
<p>Photo of the set-up of gel casting step. Line A, shows the gasket attached on the non-adhesive glass. Line B shows the deposition of the prepolymer solution inside the gasket and line C the final setup with the adhesive glass on top of the prepolymer solution.</p>
Full article ">Figure 4
<p>Example of synthetized hydrogels. Round glass coverslips are used as a rigid, transparent substrate to anchor the polymerized hydrogel.</p>
Full article ">Figure 5
<p>Elastic moduli assessed through micropipette aspiration. Experimental methods for mechanical analysis are described in a previous publication [<a href="#B23-mps-07-00075" class="html-bibr">23</a>]. Values are reported as means. Error bars represent the standard deviation.</p>
Full article ">Figure 6
<p>Mesh sizes of three different gel compositions evaluated through confocal imaging analyzing the diffusion of fluorescent dextrans with known hydrodynamic diameters (D<sub>D</sub>) inside the network of the gels. Scale bar 200 μm.</p>
Full article ">Figure 7
<p>Example of immunofluorescence results performed on U2OS cells seeded on PEG-hydrogel with controlled stiffness and adhesiveness. Nuclei are stained in blue, F-actin in red. Scale bar = 50 μm.</p>
Full article ">Figure 8
<p>Example of immunofluorescence performed on U2OS cells seeded on PEG hydrogels with controlled stiffness and adhesiveness on a glass substrate (last images on the right). Nuclei are stained in blue, F-actin in red, and YAP/TAZ in green. Scale bar = 10 μm.</p>
Full article ">Figure 9
<p>Image of gels resulting from improper deposition of solution.</p>
Full article ">
13 pages, 3979 KiB  
Article
Synthesis and Photopatterning of Synthetic Thiol-Norbornene Hydrogels
by Umu S. Jalloh, Arielle Gsell, Kirstene A. Gultian, James MacAulay, Abigail Madden, Jillian Smith, Luke Siri and Sebastián L. Vega
Gels 2024, 10(3), 164; https://doi.org/10.3390/gels10030164 - 23 Feb 2024
Viewed by 2202
Abstract
Hydrogels are a class of soft biomaterials and the material of choice for a myriad of biomedical applications due to their biocompatibility and highly tunable mechanical and biochemical properties. Specifically, light-mediated thiol-norbornene click reactions between norbornene-modified macromers and di-thiolated crosslinkers can be used [...] Read more.
Hydrogels are a class of soft biomaterials and the material of choice for a myriad of biomedical applications due to their biocompatibility and highly tunable mechanical and biochemical properties. Specifically, light-mediated thiol-norbornene click reactions between norbornene-modified macromers and di-thiolated crosslinkers can be used to form base hydrogels amenable to spatial biochemical modifications via subsequent light reactions between pendant norbornenes in the hydrogel network and thiolated peptides. Macromers derived from natural sources (e.g., hyaluronic acid, gelatin, alginate) can cause off-target cell signaling, and this has motivated the use of synthetic macromers such as poly(ethylene glycol) (PEG). In this study, commercially available 8-arm norbornene-modified PEG (PEG-Nor) macromers were reacted with di-thiolated crosslinkers (dithiothreitol, DTT) to form synthetic hydrogels. By varying the PEG-Nor weight percent or DTT concentration, hydrogels with a stiffness range of 3.3 kPa–31.3 kPa were formed. Pendant norbornene groups in these hydrogels were used for secondary reactions to either increase hydrogel stiffness (by reacting with DTT) or to tether mono-thiolated peptides to the hydrogel network. Peptide functionalization has no effect on bulk hydrogel mechanics, and this confirms that mechanical and biochemical signals can be independently controlled. Using photomasks, thiolated peptides can also be photopatterned onto base hydrogels, and mesenchymal stem cells (MSCs) attach and spread on RGD-functionalized PEG-Nor hydrogels. MSCs encapsulated in PEG-Nor hydrogels are also highly viable, demonstrating the ability of this platform to form biocompatible hydrogels for 2D and 3D cell culture with user-defined mechanical and biochemical properties. Full article
(This article belongs to the Special Issue Hydrogel-Based Scaffolds with a Focus on Medical Use (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>Mechanical characterization of PEG-Nor hydrogels. (<b>A</b>) Compressive mechanical testing is performed to measure the elastic modulus, which is calculated as the slope between 10 and 20% strain in a stress–strain curve (blue dashed box). (<b>B</b>) Elastic moduli as a function of DTT concentration for 3, 4, 5 and 6 wt% PEG-Nor hydrogel compositions. (<b>C</b>) Schematic shows an experimental design for secondary reactions of base PEG-Nor hydrogels with mono-thiolated (cRGD, cGFP) and di-thiolated molecules (DTT). (<b>D</b>) Bar graph shows elastic moduli after secondary reactions in 5 wt% PEG-Nor hydrogels with 5 and 7 mM DTT concentration. Bar graphs and scatter plot dots represent the mean and error bars represent standard deviation, * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 2
<p>Photopatterning of mono-thiolated peptides onto PEG-Nor hydrogels. (<b>A</b>) Schematic shows photopatterning process for one peptide (cRhodamine shown) and for two peptides (cRhodamine followed by cGFP). (<b>B</b>) Representative confocal image and plot profile of PEG-Nor hydrogel photopatterned with cRhodamine. (<b>C</b>) Representative confocal image and plot profile of PEG-Nor hydrogel photopatterned with cGFP. (<b>D</b>) Side view and volume view of photopatterned PEG-Nor hydrogel with cGFP. (<b>E</b>) Representative confocal image and intensity plot profiles of sequential photopatterning of PEG-Nor hydrogel with vertical cRhodamine and horizontal cGFP stripes. Scale bars: (<b>B</b>,<b>C</b>,<b>E</b>) = 100 μm.</p>
Full article ">Figure 3
<p>MSCs attach and are mechanically active on RGD-functionalized PEG-Nor hydrogels. (<b>A</b>) Schematic of the experimental design for 2D MSC PEG-Nor studies. 2D morphological analysis of cell (<b>B</b>) area, (<b>C</b>) circularity, and (<b>D</b>) aspect ratio of MSCs on RGD-functionalized PEG-Nor hydrogels formed with 5 mM or 7 mM DTT crosslinker concentrations. (<b>E</b>) Quantification of nuclear YAP localization of MSCs on RGD-functionalized PEG-Nor hydrogels formed with 5 mM or 7 mM DTT crosslinker concentrations. Representative images of single MSCs stained for cytoskeletal actin (red), nuclei (blue), and YAP (green) on top of (<b>F</b>) 5 mM DTT and (<b>G</b>) 7 mM DTT RGD-functionalized PEG-Nor hydrogels (dashed white lines denote nuclear outlines). Bars represent the mean and error bars represent standard deviation, *** <span class="html-italic">p</span> &lt; 0.001, while ns indicates not statistically significant. Scale bars: (<b>F</b>,<b>G</b>) = 25 μm.</p>
Full article ">Figure 4
<p>MSCs encapsulated in PEG-Nor hydrogels are round and highly viable. (<b>A</b>) Schematic of forming PEG-Nor hydrogels with encapsulated MSCs. 3D morphological analysis of cell (<b>B</b>) volume and (<b>C</b>) sphericity after 1, 3, and 7 days in culture. (<b>D</b>) Percentage of live MSCs after 1, 3, and 7 days in culture, and (<b>E</b>) representative confocal images of viable MSCs encapsulated in PEG-Nor hydrogels after 1, 3, and 7 days in culture. Bars represent the mean, while error bars represent standard deviation, *** <span class="html-italic">p</span> &lt; 0.001, and ns indicates not statistically significant. Scale bar: (<b>E</b>) = 100 μm.</p>
Full article ">Figure 5
<p>Synthesis of PEG-Nor hydrogels using thiol-norbornene click chemistry. (<b>A</b>) Solution containing 8-arm PEG-Nor macromer, dithiol crosslinker DTT, and photoinitiator I2959 in the presence of UV light reacts to form PEG-Nor hydrogels. (<b>B</b>) Cylindrically formed 3D PEG-Nor hydrogels with 8 mm diameter and 2 mm height.</p>
Full article ">
12 pages, 2683 KiB  
Article
Polyglycerol-Based Hydrogel as Versatile Support Matrix for 3D Multicellular Tumor Spheroid Formation
by Boonya Thongrom, Peng Tang, Smriti Arora and Rainer Haag
Gels 2023, 9(12), 938; https://doi.org/10.3390/gels9120938 - 29 Nov 2023
Cited by 1 | Viewed by 1788
Abstract
Hydrogel-based artificial scaffolds are essential for advancing cell culture models from 2D to 3D, enabling a more realistic representation of physiological conditions. These hydrogels can be customized through crosslinking to mimic the extracellular matrix. While the impact of extracellular matrix scaffolds on cell [...] Read more.
Hydrogel-based artificial scaffolds are essential for advancing cell culture models from 2D to 3D, enabling a more realistic representation of physiological conditions. These hydrogels can be customized through crosslinking to mimic the extracellular matrix. While the impact of extracellular matrix scaffolds on cell behavior is widely acknowledged, mechanosensing has become a crucial factor in regulating various cellular functions. cancer cells’ malignant properties depend on mechanical cues from their microenvironment, including factors like stiffness, shear stress, and pressure. Developing hydrogels capable of modulating stiffness holds great promise for better understanding cell behavior under distinct mechanical stress stimuli. In this study, we aim to 3D culture various cancer cell lines, including MCF-7, HT-29, HeLa, A549, BT-474, and SK-BR-3. We utilize a non-degradable hydrogel formed from alpha acrylate-functionalized dendritic polyglycerol (dPG) and thiol-functionalized 4-arm polyethylene glycol (PEG) via the thiol-Michael click reaction. Due to its high multivalent hydroxy groups and bioinert ether backbone, dPG polymer was an excellent alternative as a crosslinking hub and is highly compatible with living microorganisms. The rheological viscoelasticity of the hydrogels is tailored to achieve a mechanical stiffness of approximately 1 kPa, suitable for cell growth. Cancer cells are in situ encapsulated within these 3D network hydrogels and cultured with cell media. The grown tumor spheroids were characterized by fluorescence and confocal microscopies. The average grown size of all tumoroid types was ca. 150 µm after 25 days of incubation. Besides, the stability of a swollen gel remains constant after 2 months at physiological conditions, highlighting the nondegradable potential. The successful formation of multicellular tumor spheroids (MCTSs) for all cancer cell types demonstrates the versatility of our hydrogel platform in 3D cell growth. Full article
(This article belongs to the Special Issue Advances in Acrylate-Based Hydrogels)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>GPC chromatogram of 10 kDa dendritic polyglycerol (dPG) and illustrating picture of dPG structure.</p>
Full article ">Figure 2
<p>Synthetic routes of dPG alpha acrylate and 4-arm PEG thiol, and the schematics showing hydrogel formation.</p>
Full article ">Figure 3
<p>Hydrogel characterization: (<b>a</b>) Shear modulus graph performed by frequency sweep test at 37 °C of hydrogel sample at different incubation times; (<b>b</b>) Stiffness bar chart obtained from G′ value at 1 Hz from frequency sweep test of hydrogel sample at each incubation time; (<b>c</b>,<b>d</b>) Photographs of hydrogel prepared by using culture medium at 1 day and 14 days of incubation, respectively.</p>
Full article ">Figure 4
<p>Hydrogel characterization: (<b>a</b>) Mesh size bar chart calculated from the stiffness value of the gel sample at each incubation time; (<b>b</b>) Mass swelling ratio of hydrogel sample at different incubation times.</p>
Full article ">Figure 5
<p>Tumor spheroids growth in the hydrogel: (<b>a</b>) Brightfield images of A549 cell line cells growing from single cells on Day 1 to tumor spheroids with an average size of around 150 µm on Day 25; (<b>b</b>) Grown tumor spheroids of BT-474, HT-29, SK-BR-3, HeLa, MCF-7 and A549 cell lines on Day 25, scale bar indicates 50 µm; (<b>c</b>) Confocal image of MCTSs, cells were stained with DAPI and Phalloidin, scale bar indicates 50 µm.</p>
Full article ">Figure 6
<p>Size determination of tumor spheroids in hydrogel matrix; (<b>a</b>) Comparative size changes with all six different cell lines from within a culture period of 25 days. Data are presented as mean ± standard deviation, n = 25, <span class="html-italic">t</span>-test, shown for A549 cell line (see <a href="#app1-gels-09-00938" class="html-app">Table S1</a>), * <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, ns denotes no significance <span class="html-italic">p</span> &gt; 0.05; (<b>b</b>) Photographs of tumoroids-encapsulated hydrogel after incubation in cell medium for 2 months. The numerous white dots inside the gel are grown MCTSs whose size is visible to the eyes.</p>
Full article ">
13 pages, 1682 KiB  
Article
Multi-Armed Star-Shaped Block Copolymers of Poly(ethylene glycol)-Poly(furfuryl glycidol) as Long Circulating Nanocarriers
by Yasuhiro Nakagawa, Kotaro Ushidome, Keita Masuda, Kazunori Igarashi, Yu Matsumoto, Tatsuya Yamasoba, Yasutaka Anraku, Madoka Takai and Horacio Cabral
Polymers 2023, 15(12), 2626; https://doi.org/10.3390/polym15122626 - 9 Jun 2023
Cited by 2 | Viewed by 2002
Abstract
Multi-arm star-shaped block copolymers with precisely tuned nano-architectures are promising candidates for drug delivery. Herein, we developed 4- and 6-arm star-shaped block copolymers consisting of poly(furfuryl glycidol) (PFG) as the core-forming segments and biocompatible poly(ethylene glycol) (PEG) as the shell-forming blocks. The polymerization [...] Read more.
Multi-arm star-shaped block copolymers with precisely tuned nano-architectures are promising candidates for drug delivery. Herein, we developed 4- and 6-arm star-shaped block copolymers consisting of poly(furfuryl glycidol) (PFG) as the core-forming segments and biocompatible poly(ethylene glycol) (PEG) as the shell-forming blocks. The polymerization degree of each block was controlled by adjusting the feeding ratio of a furfuryl glycidyl ether and ethylene oxide. The size of the series of block copolymers was found to be less than 10 nm in DMF. In water, the polymers showed sizes larger than 20 nm, which can be related to the association of the polymers. The star-shaped block copolymers effectively loaded maleimide-bearing model drugs in their core-forming segment with the Diels–Alder reaction. These drugs were rapidly released upon heating via a retro Diels–Alder step. When the star-shaped block copolymers were injected intravenously in mice, they showed prolonged blood circulation, with more than 80% of the injected dose remaining in the bloodstream at 6 h after intravenous injection. These results indicate the potential of the star-shaped PFG-PEG block copolymers as long-circulating nanocarriers. Full article
(This article belongs to the Special Issue Bioactivated Polymers for Nanomedicine)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Time-dependent loading of maleimide to TTP-PFG50-PEG60 via a Diels–Alder reaction at 40 °C. (<b>b</b>) Time-dependent release of maleimide from TTP-PFG50-PEG60-pMal via a retro–Diels–Alder reaction at 80 °C.</p>
Full article ">Figure 2
<p>Cellular uptake of TTP-PFG50-PEG60-AF488 in BxPC3 cells (blue: nuclei; green: TTP-PFG50-PEG60-AF488; red: the 8-arm PEG-Cy5; yellow: co-localization of TTP-PFG50-PEG60-AF488 and the 8-arm-PEG-Cy5). Scale bar: 20 μm.</p>
Full article ">Figure 3
<p>The blood circulation of TTP-PFG50-PEG60-AF488 and the 8-arm PEG-Cy5. (<b>a</b>) IVRT-CLSM images in the earlobe of a mouse. Scale = 50 μm. (<b>b</b>) Time-dependent profile of the fluorescent intensity in the bloodstream obtained from the region of interest in (<b>a</b>).</p>
Full article ">Scheme 1
<p>Polymerization of (<b>a</b>) 4-arm PFG, and (<b>b</b>) 4-arm PFG-PEG.</p>
Full article ">Scheme 2
<p>Polymerization of (<b>a</b>) 6-arm PFG, and (<b>b</b>) 6-arm PFG-PEG.</p>
Full article ">Scheme 3
<p>Model drug loading and release via a Diels–Alder reaction. (<b>a</b>) The 4-arm P(FuGE)-PEG. (<b>b</b>) The 6-arm P(FuGE)-PEG.</p>
Full article ">
20 pages, 3723 KiB  
Article
Injectable Hydrogels Based on Cyclodextrin/Cholesterol Inclusion Complexation and Loaded with 5-Fluorouracil/Methotrexate for Breast Cancer Treatment
by Saud Almawash, Ahmed M. Mohammed, Mohamed A. El Hamd and Shaaban K. Osman
Gels 2023, 9(4), 326; https://doi.org/10.3390/gels9040326 - 12 Apr 2023
Cited by 3 | Viewed by 2106
Abstract
Breast cancer is the second most common cancer in women worldwide. Long-term treatment with conventional chemotherapy may result in severe systemic side effects. Therefore, the localized delivery of chemotherapy helps to overcome such a problem. In this article, self-assembling hydrogels were constructed via [...] Read more.
Breast cancer is the second most common cancer in women worldwide. Long-term treatment with conventional chemotherapy may result in severe systemic side effects. Therefore, the localized delivery of chemotherapy helps to overcome such a problem. In this article, self-assembling hydrogels were constructed via inclusion complexation between host β-cyclodextrin polymers (8armPEG20k-CD and pβ-CD) and the guest polymers 8-armed poly(ethylene glycol) capped either with cholesterol (8armPEG20k-chol) or adamantane (8armPEG20k-Ad) and were loaded with 5-fluorouracil (5-FU) and methotrexate (MTX). The prepared hydrogels were characterized by SEM and rheological behaviors. The in vitro release of 5-FU and MTX was studied. The cytotoxicity of our modified systems was investigated against breast tumor cells (MCF-7) using an MTT assay. Additionally, the histopathological changes in breast tissues were monitored before and after their intratumor injection. The results of rheological characterization indicated the viscoelastic behavior in all cases except for 8armPEG-Ad. In vitro release results showed a variable range of release profiles from 6 to 21 days, depending on the hydrogel composition. MTT findings indicated the inhibition ability of our systems against the viability of cancer cells depending on the kind and concentration of the hydrogel and the incubation period. Moreover, the results of histopathology showed the improvement of cancer manifestation (swelling and inflammation) after intratumor injection of loaded hydrogel systems. In conclusion, the obtained results indicated the applicability of the modified hydrogels as injectable vehicles for both loading and controlled release of anticancer therapies. Full article
(This article belongs to the Special Issue Cancer Cell Biology in Biological Hydrogel)
Show Figures

Figure 1

Figure 1
<p>Photos of self-assembling hydrogel based on CD-chol/Ad inclusion complexation: (<b>A</b>) 8armPEG20k-chol/8armPEG20k-CD, (<b>B</b>) 8armPEG20k-Ad/pβ-CD, (<b>C</b>) 8armPEG20k-chol/pβ-CD, (<b>D</b>) 8armPEG-OH/pβ-CD, and (<b>E</b>) 8armPEG20k-chol/native β-CD.</p>
Full article ">Figure 2
<p>SEM micrographs showing the channels and cross sections of different lyophilized hydrogel systems at a magnification power (×100). (<b>a</b>–<b>c</b>) The graphs display the hydrogel formulas: (<b>a</b>) formula A, (<b>b</b>) formula B, and (<b>c</b>) formula C. (<b>d</b>–<b>f</b>) The graphs display the physical mixtures of the hydrogel components: (<b>d</b>) physical mixture of formula A, (<b>e</b>) physical mixture of formula B, and (<b>f</b>) physical mixture of formula C.</p>
Full article ">Figure 3
<p>In vitro release profiles of both 5-FU and MTX from (<b>A</b>) hydrogel formula A, composed of 30%, <span class="html-italic">w</span>/<span class="html-italic">v</span> 8armPEG-chol/8rmPEG-CD (1:1%, <span class="html-italic">w</span>/<span class="html-italic">w</span> ratio), and from (<b>B</b>) hydrogel formula C, composed of 10%, <span class="html-italic">w</span>/<span class="html-italic">v</span> 8armPEG20k-chol/pβ-CD (1:1%, <span class="html-italic">w</span>/<span class="html-italic">w</span> ratio) at 37 °C in PBS.</p>
Full article ">Figure 4
<p>Cytotoxicity assay (% cell viability), as a function of drug concentration, loaded into the modified gel systems; G4 (8armPEG20k-CD/8armPEG20k-chol), G5 (pβ-CD/8armPEG20k-Ad), and G6 (pβ-CD/8armPEG20k-chol) compared with G2 (5-FU free saline solution) and G3 (5-FU/MTX free saline solution) against MCF-7 breast cancer cell line. The results are presented as the average of three independent measurements ± SD.</p>
Full article ">Figure 5
<p>Cytotoxicity assay (% of cell viability), as a function of the incubation period, of the modified gel systems; G4 (8armPEG20k-CD/8armPEG20k-chol), G5 (pβ-CD/8armPEG20k-Ad), and G6 (pβ-CD/8armPEG20k-chol) compared with G2 (5-FU free saline solution) and G3 (5-FU/MTX free saline solution) against MCF-7 breast cancer cell line. The results are presented as the average of three independent measurements ± SD.</p>
Full article ">Figure 6
<p>Photographs of the treated rats, including normal rat (<b>A</b>), untreated tumor-induced rat (<b>B</b>), and the various treated groups (<b>C</b>–<b>E</b>), including 5-FU/MTX saline solution injected group (<b>C</b>), the drug-loaded hydrogel system A (<b>D</b>), and the drug-loaded gel system C (<b>E</b>).</p>
Full article ">Figure 7
<p>Photomicrographs display some histopathological appraisals (arrows), which were detected in the mammary gland tissues from six treatment groups: (<b>A</b>) the unmedicated group showing numerous clusters of neoplastic cells with marked edema; (<b>B</b>) the group receiving (5-FU/MTX) saline solution showing proliferating neoplastic cells with inflammatory reactions; (<b>C</b>) the group receiving hydrogel formula A showing few proliferating mammary ducts and inflammatory cells infiltration; and (<b>D</b>) the group receiving hydrogel formula C showing few neoplastic cells forming ducts with mild inflammatory edema; 100× (H&amp;E).</p>
Full article ">Figure 8
<p>The percentage of relative tumor volume (%RTV) as an indication of antitumor efficacy of the modified hydrogel systems (formulas A and C) in comparison with the free drugs and untreated groups after their local injection into the breast tumor (<span class="html-italic">n</span> = 8).</p>
Full article ">Figure 9
<p>The effect of the modified hydrogel systems (A and C) loaded with dual anticancer (5-FU/MTX) on the body weight of rats in comparison with untreated animals.</p>
Full article ">
13 pages, 2627 KiB  
Article
Aggregation and Gelation Behavior of Stereocomplexed Four-Arm PLA-PEG Copolymers Containing Neutral or Cationic Linkers
by Francesca Signori, Jos W. H. Wennink, Simona Bronco, Jan Feijen, Marcel Karperien, Ranieri Bizzarri and Pieter J. Dijkstra
Int. J. Mol. Sci. 2023, 24(4), 3327; https://doi.org/10.3390/ijms24043327 - 7 Feb 2023
Cited by 3 | Viewed by 1849
Abstract
Poly(lactide) (PLA) and poly(ethylene glycol) (PEG)-based hydrogels were prepared by mixing phosphate buffer saline (PBS, pH 7.4) solutions of four-arm (PEG-PLA)2-R-(PLA-PEG)2 enantiomerically pure copolymers having the opposite chirality of the poly(lactide) blocks. Dynamic Light Scattering, rheology measurements, and fluorescence spectroscopy [...] Read more.
Poly(lactide) (PLA) and poly(ethylene glycol) (PEG)-based hydrogels were prepared by mixing phosphate buffer saline (PBS, pH 7.4) solutions of four-arm (PEG-PLA)2-R-(PLA-PEG)2 enantiomerically pure copolymers having the opposite chirality of the poly(lactide) blocks. Dynamic Light Scattering, rheology measurements, and fluorescence spectroscopy suggested that, depending on the nature of the linker R, the gelation process followed rather different mechanisms. In all cases, mixing of equimolar amounts of the enantiomeric copolymers led to micellar aggregates with a stereocomplexed PLA core and a hydrophilic PEG corona. Yet, when R was an aliphatic heptamethylene unit, temperature-dependent reversible gelation was mainly induced by entanglements of PEG chains at concentrations higher than 5 wt.%. When R was a linker containing cationic amine groups, thermo-irreversible hydrogels were promptly generated at concentrations higher than 20 wt.%. In the latter case, stereocomplexation of the PLA blocks randomly distributed in micellar aggregates is proposed as the major determinant of the gelation process. Full article
(This article belongs to the Collection Feature Papers in Molecular Nanoscience)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Synthetic route for the preparation of four-arm copolymers <b>I</b>-L, <b>II</b>-L and <b>III</b>-L. The synthesis of <b>I</b>-D, <b>II</b>-D and <b>III</b>-D is analog. See Ref. [<a href="#B45-ijms-24-03327" class="html-bibr">45</a>] for details. (<b>b</b>) Example of gelation in PBS upon mixing 20 wt.% solutions of enantiomerically pure copolymers <b>II</b>-L and <b>II</b>-D.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) DLS intensity plots of sc-<b>I</b> (<b>a</b>), sc-<b>II</b> (<b>b</b>), and sc-<b>III</b> (<b>c</b>) immediately after mixing (solid line) and after 18 h (dashed line) at 37 °C. (<b>d</b>–<b>f</b>) Relative DLS intensity (range 10–100 nm) plots after mixing equimolar solutions of copolymers <b>I</b>-L + <b>I</b>-D (<b>d</b>), <b>II</b>-L + <b>II</b>-D (<b>e</b>) and <b>III</b>-L + <b>III</b>-D (<b>f</b>) at 25 °C (blue circles), 37 °C (green circles), and 50 °C (red circles); solid lines represent linear (<b>d</b>) or exponential (<b>e</b>,<b>f</b>) fits to datasets. All experiments were carried out in PBS using 0.3 wt.% of L and D copolymers.</p>
Full article ">Figure 3
<p>Fluorescence emission (left axis) and anisotropy (right axis) of Ge1 in the 440–700 interval upon 420 nm excitation. Red: <b>II</b>-L. Blue: sc-<b>II</b> immediately after mixing. Black: sc-<b>II</b> 72 h after mixing. Experiments were carried out in PBS at 25 °C using 0.3 wt.% of L and D copolymers.</p>
Full article ">Figure 4
<p>Temperature-dependent storage (G′) and loss (G″) moduli of sc-<b>I</b> (<b>a</b>) and sc-<b>II</b> (<b>b</b>) hydrogels upon heating to 75 °C and subsequent cooling to 25 °C (solid grey line). Arrows in (<b>a</b>) indicate the sol-gel and gel-sol transitions. In both cases the solutions were prepared in PBS, pH = 7.4. Concentrations: sc-<b>I,</b> 10 wt.%; sc-<b>II</b>, 20 wt.%.</p>
Full article ">Figure 5
<p>Model of hydrogel formation for block copolymers <b>I</b>-L <b>and I</b>-D (<b>a</b>) and block copolymers <b>II</b>-L and <b>II</b>-D and <b>III</b>-L and <b>III</b>-D (<b>b</b>).</p>
Full article ">
11 pages, 3269 KiB  
Article
Study of Excipients in Delayed Skin Reactions to mRNA Vaccines: Positive Delayed Intradermal Reactions to Polyethylene Glycol Provide New Insights for COVID-19 Arm
by David Pesqué, Ramon Maria Pujol, Orianna Marcantonio, Ainhoa Vidal-Navarro, José María Ramada, Alba Arderiu-Formentí, Agustí Albalat-Torres, Consol Serra and Ana María Giménez-Arnau
Vaccines 2022, 10(12), 2048; https://doi.org/10.3390/vaccines10122048 - 30 Nov 2022
Cited by 1 | Viewed by 3268
Abstract
Background: Skin local reactions to mRNA COVID-19 vaccines have been linked to the use of vaccine excipients. The aim of the study is to evaluate the role of skin testing excipients in delayed skin reactions due to mRNA COVID-19 vaccines. Methods: Skin testing [...] Read more.
Background: Skin local reactions to mRNA COVID-19 vaccines have been linked to the use of vaccine excipients. The aim of the study is to evaluate the role of skin testing excipients in delayed skin reactions due to mRNA COVID-19 vaccines. Methods: Skin testing among a group of healthcare workers with skin reactions due to mRNA vaccines was performed. Patch testing and intradermal testing (IDT) with polyethylene glycol (PEG)-400, PEG-2000, trometamol, and 1,2-dimyristoyl-sn-glycero-3-phosphocholine were performed. Healthcare workers without skin reactions to vaccines were used for skin testing as controls. Results: Thirty-one healthcare workers (from a total of 4315 vaccinated healthcare workers) experienced cutaneous adverse vaccine reactions. Skin testing was performed in sixteen of the healthcare workers (11 delayed large local reactions (DLLR) and 5 widespread reactions). Positive IDT for PEG-2000 1% in DLLR was seen in 10 (90.9%) patients, in comparison with one (16.6%) individual with a delayed widespread reaction. Delayed positive IDT reactions for PEG-2000 1% on day 2 were observed in three (27.3%) patients with DLLR. Patch testing of the excipients was negative. Among 10 controls, only one exhibited a transient positive IDT reaction to PEG-2000 1%. Conclusions: Immediate and delayed reactions to IDT are frequently detected in patients with DLLR. The observation of positive delayed intradermal reactions to PEG disclosed only in patients with DLLR reinforces a possible role of PEG in the development of these reactions. Skin testing of other excipients is of little importance in clinical practice. Full article
(This article belongs to the Special Issue Adverse Events of COVID-19 Vaccines)
Show Figures

Figure 1

Figure 1
<p>Skin testing: results for polyethylene glycol. (<b>A</b>) Intradermal testing (IDT) of PEG-2000 with positivity for all concentrations with significant papules and erythema at 20 min. (<b>B</b>) IDT of PEG-2000 with positivity for PEG-2000 1%, 0.1%, and 0.01% at 2 h.</p>
Full article ">
13 pages, 4375 KiB  
Article
Facile Construction of Hybrid Hydrogels with High Strength and Biocompatibility for Cranial Bone Regeneration
by Shuai Chang, Jiedong Wang, Nanfang Xu, Shaobo Wang, Hong Cai, Zhongjun Liu and Xing Wang
Gels 2022, 8(11), 745; https://doi.org/10.3390/gels8110745 - 17 Nov 2022
Cited by 6 | Viewed by 2413
Abstract
The significant efforts being made towards the utilization of artificial soft materials holds considerable promise for developing tissue engineering scaffolds for bone-related diseases in clinics. However, most of these biomaterials cannot simultaneously satisfy the multiple requirements of high mechanics, good compatibility, and biological [...] Read more.
The significant efforts being made towards the utilization of artificial soft materials holds considerable promise for developing tissue engineering scaffolds for bone-related diseases in clinics. However, most of these biomaterials cannot simultaneously satisfy the multiple requirements of high mechanics, good compatibility, and biological osteogenesis. In this study, an osteogenic hybrid hydrogel between the amine-functionalized bioactive glass (ABG) and 4-armed poly(ethylene glycol) succinimidyl glutarate-gelatin network (SGgel) is introduced to flexibly adhere onto the defective tissue and to subsequently guide bone regeneration. Relying on the rapid ammonolysis reaction between amine groups (-NH2) of gelatin and ABG components and N-hydroxysuccinimide (NHS)-ester of tetra-PEG-SG polymer, the hydrogel networks were formed within seconds, offering a multifunctional performance, including easy injection, favorable biocompatibility, biological and mechanical properties (compressive strength: 4.2 MPa; storage modulus: 104 kPa; adhesive strength: 56 kPa), which could facilitate the stem cell viability, proliferation, migration and differentiation into osteocytes. In addition, the integration between the SGgel network and ABG moieties within a nano-scale level enabled the hybrid hydrogel to form adhesion to tissue, maintain the durable osteogenesis and accelerate bone regeneration. Therefore, a robust approach to the simultaneously satisfying tough adhesion onto the tissue defects and high efficiency for bone regeneration on a mouse skull was achieved, which may represent a promising strategy to design therapeutic scaffolds for tissue engineering in clinical applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the fabrication procedures of SGgel@ABG composite hydrogel for calvaria bone defects repair.</p>
Full article ">Figure 2
<p>Structure and property characterizations. (<b>A</b>) Synthesis route of modified polymers and amine-functionalized ABG. (<b>B</b>) <sup>1</sup>H NMR spectrum of the tetra-PEG-SG polymer. (<b>C</b>–<b>F</b>) SEM images, compressive, rheology and adhesive profiles of (<b>a</b>) SGgel and (<b>b</b>) SGgel@ABG hydrogels. Red arrows represent the similar inner pores.</p>
Full article ">Figure 3
<p>Cell cytotoxicity of SGgel and SGgel@ABG hybrid scaffolds in vitro. (<b>A</b>) Live/dead staining of BMSCs. Wherein, the green cells are the living BMSCs, and the red cells are the dead BMSCs. (<b>B</b>) Cell viability and (<b>C</b>) Cell proliferation of SGgel and SGgel@ABG hybrid scaffolds after the cultivation for the appointed time. NS, not significant.</p>
Full article ">Figure 4
<p>In vitro osteogenic differentiation of SGgel@ABG hydrogel. (<b>A</b>–<b>C</b>) ALP (14 d) and ARS staining (21 d) revealing the enhanced osteogenic differentiation of BMSCs. (<b>D</b>,<b>E</b>) Western blotting analysis and (<b>F</b>) qPCR quantification showing the highest osteogenic expression markers (OCN, ALP, Osterix and RUNX2) in the hydrogels. Statistically significant differences in comparison with untreated cells (control), SGgel hydrogel and SGgel@ABG hydrogel. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>(<b>A</b>) 3D reconstruction of Micro-CT images of regenerated bone formation in rat cranium after the hydrogel implantation for 8 weeks with control group, SGgel group and SGgel@ABG group. (<b>B</b>–<b>D</b>) Quantitative analysis of BV, BV/TV and BMD of newly formed bone tissue. (<b>E</b>) H&amp;E and Masson’s trichrome staining. (<b>F</b>,<b>G</b>) Woven bone and cartilage areas were analyzed in defect bone region. ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">
12 pages, 4968 KiB  
Article
Hydrogel Dressing Containing Basic Fibroblast Growth Factor Accelerating Chronic Wound Healing in Aged Mouse Model
by Yonghao Xiao, Hui Zhao, Xiaoyu Ma, Zongheng Gu, Xin Wu, Liang Zhao, Lin Ye and Zengguo Feng
Molecules 2022, 27(19), 6361; https://doi.org/10.3390/molecules27196361 - 26 Sep 2022
Cited by 16 | Viewed by 2742
Abstract
Due to the decreasing self-repairing ability, elder people are easier to form chronic wounds and suffer from slow and difficult wound healing. It is desirable to develop a novel wound dressing that can accelerate chronic wound healing in elderly subjects to decrease the [...] Read more.
Due to the decreasing self-repairing ability, elder people are easier to form chronic wounds and suffer from slow and difficult wound healing. It is desirable to develop a novel wound dressing that can accelerate chronic wound healing in elderly subjects to decrease the pain of patients and save medical resources. In this work, Heparin and basic fibroblast growth factor(bFGF) were dissolved in the mixing solution of 4-arm acrylated polyethylene glycol and dithiothreitol to form hydrogel dressing in vitro at room temperature without any catalysts, which is convenient and easy to handle in clinic application. In vitro re-lease test shows the bFGF could be continuously released for at least 7 days, whereas the dressing surface integrity maintained for 3 days degradation in PBS solution. Three groups of treatments including bFGF-Gel, bFGF-Sol and control without any treatment were applied on the full-thickness wound on the 22 months old mice back. The wound closure rate and histological and immunohistochemical staining all illustrated that bFGF-Gel displayed a better wound healing effect than the other two groups. Thus, as-prepared hydrogel dressing seems supe-rior to current clinical treatment and more effective in elderly subjects, which shows promising potential to be applied in the clinic. Full article
(This article belongs to the Special Issue Polymer Scaffolds for Biomedical Applications III)
Show Figures

Figure 1

Figure 1
<p>The profiles of bFGF-Gel in vitro. (<b>A</b>) The swelling behavior of bFGF-Gel. (<b>B</b>) bFGF release of the bFGF-Gel. (<b>C</b>) Storage modulus G′ and loss modulus G″ versus time during the bFGF-Gel degradation process.</p>
Full article ">Figure 2
<p>The different morphology of bFGF-Gel degraded after 0 days (<b>A</b>), 3 days (<b>B</b>), 6 days (<b>C</b>), 9 days (<b>D</b>), 12 days (<b>E</b>), 15 days (<b>F</b>) observed by SEM. The length of the scale bar is 100 μm.</p>
Full article ">Figure 3
<p>The evaluation of chronic wound healing on the aged mouse. (<b>A</b>) Full-thickness skin wound of bFGF-Gel group, (<b>B</b>) bFGF-Gel dressing covered on the wound, (<b>C</b>) Full-thickness skin wound of bFGF-Sol group, (<b>D</b>) bFGF was spraying on the wound, (<b>E</b>) Digital photos of wounds on the dorsum of mice after treatment by using no treatment (Blank), spraying bFGF-Sol and bFGF-Gel hydrogel dressings on different days after the surgery. (<b>F</b>) Summary graph of the wound closure rate of these three groups. The data were represented as mean ± SD (<span class="html-italic">n</span> = 3). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001. Scale bar =10 mm.</p>
Full article ">Figure 4
<p>Histological analysis of wounds on the dorsum of mice by H&amp;E staining after treatment with no bFGF, spraying bFGF, and hydrogel dressing containing bFGF for up to 10 days. (<b>A</b>) normal skin, (<b>B</b>) Blank, (<b>C</b>) bFGF-Sol, (<b>D</b>) bFGF-Gel. Scale bar =100 μm. (Green arrows: fibroblasts; black arrows: collagenous fibers; Red arrows: blood capillaries.). Masson staining images from day 10 post-operation. (<b>E</b>) normal skin, (<b>F</b>) Blank, (<b>G</b>) bFGF, (H) bFGF-Gel, (<b>I</b>) relative content of collagen in different groups compared with normal skin. Scale bar =100 μm.</p>
Full article ">Figure 5
<p>Immunohistochemical analyses of wounds on the dorsum of mice by CD31, CD68, and Ki67 staining after 10 days. (<b>A</b>) Representative images of CD31, CD68, and Ki67 staining in groups of Normal skin, Blank, bFDF, and bFGF-Gel at day 10; Scale bar = 200 μm. (<b>B</b>) Semiquantitative results of CD31, CD68, and Ki67 staining. The data were represented as mean ± SD (<span class="html-italic">n</span> =3). * <span class="html-italic">p</span> &lt; 0.05, *** <span class="html-italic">p</span> &lt; 0.001. Scale bar = 200 μm.</p>
Full article ">
13 pages, 3928 KiB  
Article
Metabolic Study of Tetra-PEG-Based Hydrogel after Pelvic Implantation in Rats
by Baoyan Zuo, Mingxue Cao, Xiumei Tao, Xiaoyu Xu, Hongfei Leng, Yali Cui and Kaishun Bi
Molecules 2022, 27(18), 5993; https://doi.org/10.3390/molecules27185993 - 14 Sep 2022
Cited by 3 | Viewed by 2282
Abstract
In vivo metabolism of polyethylene glycol (PEG) hydrogels has rarely been studied. In this study, we prepared a chemically crosslinked hydrogel formulation using 14C-labeled tetra-armed poly (ethylene glycol) succinimidyl succinate (Tetra-PEG-SS) and 3H-labeled crosslinking agent for implantation into the pelvis of [...] Read more.
In vivo metabolism of polyethylene glycol (PEG) hydrogels has rarely been studied. In this study, we prepared a chemically crosslinked hydrogel formulation using 14C-labeled tetra-armed poly (ethylene glycol) succinimidyl succinate (Tetra-PEG-SS) and 3H-labeled crosslinking agent for implantation into the pelvis of Sprague-Dawley (SD) rats. This radioactive labeling technique was used to investigate the radioactivity excretion rates in of feces and urine, the blood exposure time curve, and the radioactivity recovery rate in each tissue over time. We showed that the primary excretion route of the hydrogel was via urine (3H: about 86.4%, 14C: about 90.0%), with fewer portion through feces (3H: about 6.922%, 14C: about 8.16%). The hydrogel metabolites exhibited the highest distribution in the kidney, followed by the jejunal contents; The 3H and 14C radioactivity exposures in the remaining tissues were low. We also showed that the 3H and 14C radioactivity recovery rates in the blood were usually low (<0.10% g−1 at 12 h after implantation), even though, in theory, the hydrogel could be absorbed into the blood through the adjacent tissues. By using a combination of HPLC-MS/MS and offline radioactivity counting method, we established that the tetra-PEG-based hydrogel was mainly metabolized to lower-order PEG polymers and other low-molecular-weight substances in vivo. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Schematic diagram of hydrogel formation. (<b>B</b>) Schematic diagram of tetra-PEG-SS radiolabeling (*).</p>
Full article ">Figure 2
<p>Concentration–time curve of average radioactivity recovery rates in the blood (mean ± SD, <span class="html-italic">n</span> = 6). (Student–Newman–Keulsa).</p>
Full article ">Figure 3
<p>Concentration–time changes of <sup>3</sup>H (<b>a</b>) and <sup>14</sup>C (<b>b</b>) radioactivity in different tissues (mean ± SD, <span class="html-italic">n</span>=6). (Student–Newman–Keulsa).</p>
Full article ">Figure 4
<p><sup>3</sup>H (<b>a</b>) and <sup>14</sup>C (<b>b</b>) radio-chromatograms of blank and 12 h urine.</p>
Full article ">Figure 5
<p>Primary mass spectrum of the control and the experimental group urine samples at the 12 h time point within RT 8.0–10.0 min.</p>
Full article ">Figure 6
<p>Total ion flow chromatogram of the control and the experimental group urine samples at the 12 h time point.</p>
Full article ">Figure 7
<p>Primary mass spectrum of <span class="html-italic">m</span>/<span class="html-italic">z</span> 400–500 extracted from the control and the experimental group urine at 12 h time point within RT 8.0–10.0 min.</p>
Full article ">Figure 8
<p>Chromatograms of <span class="html-italic">m/z</span> 415.25 were extracted from the control and the experimental group urine samples at the 12 h time point.</p>
Full article ">
Back to TopTop