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

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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (10,064)

Search Parameters:
Keywords = scaffold

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 5582 KiB  
Article
Comparison of Two Chelator Scaffolds as Basis for Cholecystokinin-2 Receptor Targeting Bimodal Imaging Probes
by Giacomo Gariglio, Katerina Bendova, Martin Hermann, Asta Olafsdottir, Jane K. Sosabowski, Milos Petrik, Elisabeth von Guggenberg and Clemens Decristoforo
Pharmaceuticals 2024, 17(12), 1569; https://doi.org/10.3390/ph17121569 - 22 Nov 2024
Abstract
Background/Objectives: Dual-modality probes, combining positron emission tomography (PET) with fluorescence imaging (FI) capabilities in a single molecule, are of high relevance for the accurate staging and guided resection of tumours. We herein present a pair of candidates targeting the cholecystokinin-2 receptor (CCK2R), [...] Read more.
Background/Objectives: Dual-modality probes, combining positron emission tomography (PET) with fluorescence imaging (FI) capabilities in a single molecule, are of high relevance for the accurate staging and guided resection of tumours. We herein present a pair of candidates targeting the cholecystokinin-2 receptor (CCK2R), namely [68Ga]Ga-CyTMG and [68Ga]Ga-CyFMG. In these probes, the SulfoCy5.5 fluorophore and two units of a CCK2R-binding motif are coupled to the chelator acting as a core scaffold, triazacyclononane-phosphinic acid (TRAP), and Fusarinine C (FSC), respectively. Using this approach, we investigated the influence of these chelators on the final properties. Methods: The synthetic strategy to both precursors was based on the stoichiometric conjugation of the components via click chemistry. The characterization in vitro included the evaluation of the CCK2R affinity and internalization in A431-CCK2R cells. Ex vivo biodistribution as well as PET and FI studies were performed in xenografted mice. Results: 68Ga labelling was accomplished with high radiochemical yield and purity for both precursors. A CCK2R affinity in the subnanomolar range of the conjugates and a receptor-specific uptake of the radioligands in cells were observed. In A431-CCK2R/A431-mock xenografted mice, the investigated compounds showed specific accumulation in the tumours and reduced off-target uptake compared to a previously developed compound. Higher accumulation and prolonged retention in the kidneys were observed for [68Ga]Ga-CyTMG when compared to [68Ga]Ga-CyFMG. Conclusions: Despite the promising targeting properties observed, further probe optimization is required to achieve enhanced imaging contrast at early timepoints. Additionally, the results indicate a distinct influence of the chelators in terms of renal accumulation and retention. Full article
(This article belongs to the Special Issue Development of Novel Radiopharmaceuticals for SPECT and PET Imaging)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of conjugates. Structure of CyTMG (<b>a</b>) and CyFMG (<b>b</b>).</p>
Full article ">Figure 2
<p>Cell-associated radioactivity determined for [<sup>68</sup>Ga]Ga-CyTMG and [<sup>68</sup>Ga]Ga-CyFMG in A431-CCK2R and A431-mock cells. The values are reported as means of three independent experiments.</p>
Full article ">Figure 3
<p>Cell-associated fluorescence determined for CyTMG and CyFMG using A431-CCK2R and A431-mock cells. The images reported overlay the SulfoCy5.5 channel (green) showing the fluorescent agent, the WGA-Alexa fluor channel (red) showing the cell membrane, and the HOECHST channel (blue) showing the nucleus. Magnification is equal for all images (scale bar: 16 µm).</p>
Full article ">Figure 4
<p>Receptor-binding affinity of CyTMG and CyFMG on A431-CCK2R cells. IC50 values are reported as means of two independent experiments, each performed in triplicate.</p>
Full article ">Figure 5
<p>Ex vivo biodistribution studies in healthy BALB/C mice (n = 3) performed at 1, 2, and 4 h p.i. for [<sup>68</sup>Ga]Ga-CyTMG and for [<sup>68</sup>Ga]Ga-CyFMG (amount injected: 0.14 nmol, 1.5 MBq). The asterisks represent the level of significance determined by using the <span class="html-italic">p</span> value (*: 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05; **: 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 6
<p>Ex vivo biodistribution studies in A431-CCK2R and A431-mock xenografted mice. BALB/C mice (n = 3) performed for 2 h p.i. for [<sup>68</sup>Ga]Ga-CyTMG and [<sup>68</sup>Ga]Ga-CyFMG (amount injected: 0.05 nmol, 1.0 MBq). The asterisks represent the level of significance determined by using the <span class="html-italic">p</span> value (*: 0.01 &lt; <span class="html-italic">p</span> &lt; 0.05; **: 0.001 &lt; <span class="html-italic">p</span> &lt; 0.01; ***: <span class="html-italic">p</span> &lt; 0.001).</p>
Full article ">Figure 7
<p>Tumour-to-organ ratios for A431 xenografts of mice injected with [<sup>68</sup>Ga]Ga-CyTMG and [<sup>68</sup>Ga]Ga-CyFMG.</p>
Full article ">Figure 8
<p>(<b>a</b>) Imaging study of two A431-CCK2R and A431-mock xenografted mice injected with [<sup>68</sup>Ga]Ga-CyTMG or [<sup>68</sup>Ga]Ga-CyFMG (amount injected: 0.3 nmol, 6.0 MBq). Static PET/CT MIP images (left) and corresponding near-infrared fluorescence images (right) at various timepoints. B: bladder; H: heart; L: liver; T: tumour; K: kidneys. (<b>b</b>) Comparison of uptake in A431-CCK2R and A431-mock tumours of animals injected with [<sup>68</sup>Ga]Ga-CyTMG or [<sup>68</sup>Ga]Ga-CyFMG. Values are calculated in manually drawn ROIs fitting to the tumours. Results are expressed as mean uptake values (n = 3).</p>
Full article ">Figure 9
<p>Cryo-fluorescence tomography of two A431-CCK2R and A431-mock xenografted mice injected with [<sup>68</sup>Ga]Ga-CyTMG or [<sup>68</sup>Ga]Ga-CyFMG performed for 72 h p.i.</p>
Full article ">
40 pages, 15332 KiB  
Review
Photocrosslinkable Biomaterials for 3D Bioprinting: Mechanisms, Recent Advances, and Future Prospects
by Yushang Lai, Xiong Xiao, Ziwei Huang, Hongying Duan, Liping Yang, Yuchu Yang, Chenxi Li and Li Feng
Int. J. Mol. Sci. 2024, 25(23), 12567; https://doi.org/10.3390/ijms252312567 - 22 Nov 2024
Abstract
Constructing scaffolds with the desired structures and functions is one of the main goals of tissue engineering. Three-dimensional (3D) bioprinting is a promising technology that enables the personalized fabrication of devices with regulated biological and mechanical characteristics similar to natural tissues/organs. To date, [...] Read more.
Constructing scaffolds with the desired structures and functions is one of the main goals of tissue engineering. Three-dimensional (3D) bioprinting is a promising technology that enables the personalized fabrication of devices with regulated biological and mechanical characteristics similar to natural tissues/organs. To date, 3D bioprinting has been widely explored for biomedical applications like tissue engineering, drug delivery, drug screening, and in vitro disease model construction. Among different bioinks, photocrosslinkable bioinks have emerged as a powerful choice for the advanced fabrication of 3D devices, with fast crosslinking speed, high resolution, and great print fidelity. The photocrosslinkable biomaterials used for light-based 3D printing play a pivotal role in the fabrication of functional constructs. Herein, this review outlines the general 3D bioprinting approaches related to photocrosslinkable biomaterials, including extrusion-based printing, inkjet printing, stereolithography printing, and laser-assisted printing. Further, the mechanisms, advantages, and limitations of photopolymerization and photoinitiators are discussed. Next, recent advances in natural and synthetic photocrosslinkable biomaterials used for 3D bioprinting are highlighted. Finally, the challenges and future perspectives of photocrosslinkable bioinks and bioprinting approaches are envisaged. Full article
(This article belongs to the Special Issue Bioprinting: Progress and Challenges)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of (<b>a</b>) extrusion, (<b>b</b>) inkjet, (<b>c</b>) stereolithography, and (<b>d</b>) light-assisted 3D bioprinting (The figures were created with BioRender).</p>
Full article ">Figure 2
<p>A schematic illustration of the mechanism of free radical chain-growth polymerization. (<b>a</b>) The mechanism of free radical chain-growth polymerization. (<b>b</b>) A schematic of polymer chains containing reactive groups crosslinking through free radical chain-growth polymerization.</p>
Full article ">Figure 3
<p>Schematic illustrations of the mechanism of thiol-ene-mediated polymerization. (<b>a</b>) A schematic illustration of the mechanism of thiol-ene crosslinking. (<b>b</b>) A schematic of polymer chains containing reactive groups crosslinking through thiol-ene polymerization. (<b>c</b>) Kinetic modeling of photoinitiated thiol-ene click chemistry based on alkene conversion and a summary of the reactivity of the alkene group (Copyright 2012 American Chemical Society [<a href="#B122-ijms-25-12567" class="html-bibr">122</a>]).</p>
Full article ">Figure 4
<p>The mechanism of redox crosslinking. (<b>a</b>) A schematic illustration of the mechanism of redox polymerization. (<b>b</b>) A schematic of polymer chains containing reactive groups crosslinking through redox reactions.</p>
Full article ">Figure 5
<p>Collagen-based photocrosslinking bioinks. (<b>a</b>) Schematic representation of methacrylate- [<a href="#B171-ijms-25-12567" class="html-bibr">171</a>,<a href="#B172-ijms-25-12567" class="html-bibr">172</a>], maleic- [<a href="#B173-ijms-25-12567" class="html-bibr">173</a>], norbornene- [<a href="#B170-ijms-25-12567" class="html-bibr">170</a>], and thiol-modified [<a href="#B174-ijms-25-12567" class="html-bibr">174</a>] collagen synthesis. (<b>b</b>) The relative solubility of NorCol and collagen at different pH values. (<b>c</b>) The miscibility of NorCol with gelatin and alginate. (<b>d</b>) Temperature-sensitive extrusion bioprinting of NorCol bioinks. (<b>d</b>) (i) Schematic of temperature-sensitive extrusion bioprinting of NorCol bio-inks. (ii) Printed NorCol hydrogels (12 layers, 3 mm) after 1 day of culture. Fluorescence micrographs showing cell (iii) viability (day 1) and (iv) spreading (day 5) within NorCol hydrogels (Copyright 2021 American Chemical Society [<a href="#B170-ijms-25-12567" class="html-bibr">170</a>]).</p>
Full article ">Figure 6
<p>Schematic illustration of collagen hydrolysis and representative routes to synthesize methacrylate, norbornene, vinyl, and tyrosine-modified gelatin.</p>
Full article ">Figure 7
<p>The bioprinting performance of photocrosslinkable gelatin bioinks. (<b>a</b>) The cell viability within the printed scaffold is affected by the methacrylate degree of GelMA. (i) Photocrosslinking for solidification. (ii) Evaluation of live and dead cells encapsulating in 7.5% GM-30/60/90 on day 5. (iii) Semiquantitative analysis of cell viability, (<span class="html-italic">** p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001) (Copyright 2023 Wiley [<a href="#B52-ijms-25-12567" class="html-bibr">52</a>]). (<b>b</b>) An illustrative scheme of cell-laden bioprinting using GelMA and GelNB/HepSH bioinks. (<b>c</b>) HUVEC-laden constructs are built from GelNB/HepSH and GelMA bioinks. (i) Fluorescence micrographs showing the bioprinted constructs after 1 and 7 days of culture. (ii) Semiquantitative analysis of cell viability. Fluorescence micrographs showing HUVEC cytoskeleton in both bioinks after cell culture for 7 days, using (iii) an inverted fluorescence microscope and (iv) a laser scanning confocal microscope, scale bars = 500 μm. (v,vi) 3D-printed canine peripheral-nerve-like constructs using the GelNB/HepSH bioink, scale bars = 4 mm (Copyright 2021 American Chemical Society [<a href="#B112-ijms-25-12567" class="html-bibr">112</a>]). (<b>d</b>) GelNB/GelS bioinks can undergo superfast gelation at extremely low photoinitiator concentrations. (i) Water-based synthesis of GelNB and GelS from gelatin. (ii) Photocrosslinked thiol-ene click hydrogel. (iii) Comparison of the two thiol-ene hydrogel systems GelNB/DTT and GelNB/GelS. <span class="html-italic">*** p</span> &lt; 0.001. (<b>e</b>) The 3D bioprinting of an NHDF-laden hydrogel grid structure (i) 3D bioprinting of a hydrogel grid structure (1 cm × 1 cm) consisting of four layers on a glass slide. Post-printing cell viability analysis of 3D bioprinted NHDF at day 1 using (ii) GelMA and (iii) GelNB/GelS bioinks. (iv) Distribution of NHDF within the hydrogel, (i) Live/dead staining and (ii) distribution of NHDF, scale bars = 100 μm, (<span class="html-italic">** p &lt;</span> 0.01) (Copyright 2021 Wiley [<a href="#B180-ijms-25-12567" class="html-bibr">180</a>]). (<b>f</b>) A schematic comparison of GelNB synthesized by (i) 5-norbornene-2-carboxylic acid and (ii) CA.</p>
Full article ">Figure 8
<p>Hyaluronic acid-based photocrosslinking bioinks. (<b>a</b>) Synthesis route of light-cured hyaluronic acid. (<b>b</b>) Modification of sodium HA with CA to form NorHA<sub>CA</sub>. (i) Reaction scheme for NorHA<sub>CA</sub> synthesis. (ii) Degree of modification of HA with norbornene is tuned by changing the molar ratio of CA to HA repeat units. (iii) Schematic representation of network formation by visible light-induced thiol-ene step-growth reaction between NorHA<sub>CA</sub> and DTT in the presence of photoinitiator (LAP), (<span class="html-italic">** p &lt;</span> 0.01, <span class="html-italic">**** p &lt;</span> 0.0001). (<b>c</b>) Biocompatibility and DLP-based 3D bioprinting of NorHA<sub>CA</sub> bioinks. (i) Representative fluorescence micrographs of bMSCs encapsulated in NorHA<sub>CA</sub> (5wt%, 40%mod.) bulk hydrogels over time (1, 3, and 7 days), scale bars = 200 μm. (ii) Semiquantitative analysis of cell viability, (<span class="html-italic">* p &lt;</span> 0.05). (iii) Schematic representation of DLP-based 3D printing of NorHACA hydrogels with bMSCs. (iv) Representative maximum projection image of bMSCs encapsulated in a NorHA<sub>CA</sub> macroporous lattice at day 1, scale bars = 1 mm and 500 μm (Copyright 2023 American Chemical Society [<a href="#B204-ijms-25-12567" class="html-bibr">204</a>]). (<b>d</b>) HA-TBA mediates the synthesis of NorHA.</p>
Full article ">Figure 9
<p>Alginate-based bioinks. (<b>a</b>) The (i) structure and (ii) gelation mechanism of alginate. (<b>b</b>) The norbornene alginate bioink (Alg-norb) was functionalized by light-mediated RGD grafting for building L929 cell-embedded constructs. (i) Schematic overview of the strategy employed to develop photoactive Alg-norb for bioprinting. (ii) Photoinitiated thiol−ene reactions of Alg-norb with RGD Peptide Sequence (CGGGRGDS). (iii) Images of 3D bioprinted hydrogels loaded with cells at (a) day 0 and (b) day 7. Green and red cell tracker labeled L929 as two different bioinks printed as alternating fibers (c) in the X-Y plane and (d) in the Z direction (Copyright 2018 American Chemical Society [<a href="#B218-ijms-25-12567" class="html-bibr">218</a>]). (<b>c</b>) The synthesis of Alg-RGD through a thiol-ene click reaction to promote the cell growth and vascularization of HUVECs. (i) Design of the HA/Alg-RGD hydrogel. (ii) Schematic diagram of the 3D printing process. (iii) Fluorescent images of GFP-HUVECs cultured in the hydrogel at intervals of 3, 7, and 14 days post-3D printing, along with magnified images of selected regions (scale bar = 1 mm and 200 μm, respectively) (Copyright 2023 American Chemical Society [<a href="#B219-ijms-25-12567" class="html-bibr">219</a>]).</p>
Full article ">Figure 10
<p>SF-based photocrosslinkable bioink. (<b>a</b>) Production, structure, and modification of SF (Copyright 2023 Elsevier [<a href="#B228-ijms-25-12567" class="html-bibr">228</a>]). (<b>b</b>) The design and printing performance of the redox-crosslinkable SF/CG bioink. (i) Schematic of the printing process of SF/CG bioink. (ii) CAD images depicting the ear, nose, and hand and printed images at various angles (Copyright 2024 Elsevier [<a href="#B237-ijms-25-12567" class="html-bibr">237</a>]). (<b>c</b>) Methacrylate-group-functionalized SF for recapitulating human skin models through 3D bioprinting (Copyright 2024 Wiely [<a href="#B243-ijms-25-12567" class="html-bibr">243</a>]).</p>
Full article ">Figure 11
<p>dECM-based photocrosslinkable bioinks. (<b>a</b>) The production and modification of dECM (Copyright 2023 Ivyspring [<a href="#B252-ijms-25-12567" class="html-bibr">252</a>]). (<b>b</b>) The Ru/SPS-induced visible light crosslinking of dECM. (i) The crosslinking mechanism. (ii) The gelled dECM hydrogel (Copyright 2023 Wiely [<a href="#B253-ijms-25-12567" class="html-bibr">253</a>]). (<b>c</b>) The light-activated dityrosine crosslinking reaction in dECM bioink to realize centimeter-scale 3D bioprinting. (i) A schematic of visible-light active dityrosine synthesis. (ii) Extrusion-based printing of dECM. (iii) DLP photopatterning with 100 µm step-size constructs, scale bars = 100 μm for white represent printed fiber, 500 μm for live/dead images, (Copyright 2021 Wiely [<a href="#B124-ijms-25-12567" class="html-bibr">124</a>]). (<b>d</b>) Liver dECM was functionalized by glycidyl methacrylate and methacrylic anhydride for the systematic comparison of different type of methacrylate dECM bioinks. The (i) preparation and (ii) modification of live dECM (Copyright 2024 Elsevier [<a href="#B254-ijms-25-12567" class="html-bibr">254</a>]). (<b>e</b>) The decellularized small intestine submucosa (dSIS) was functionalized by norbornene to create an orthogonally crosslinked dSIS hydrogel for cancer and vascular tissue engineering. (i) 1H NMR spectra of dSIS and dSIS-NB. Peak a: alkene protons (HC=CH), Peak b: ethyl protons (CH<sub>2</sub>), Peaks c and d: methine protons (C<sub>3</sub>CH). (ii) Schematic of thiol-norbornene photo-crosslinking. (iii) Schematic of DLP bioprinting. (iv) In situ, photo rheometry of dSIS-NB gelation with tartrazine added as a photo absorber to improve printing fidelity, dotted line indicate light on. (v) A CAD image of astar-shaped object for DLP bioprinting and the DLP printed dSIS-NB gel. (vi) A representative live/dead confocal image of interconnected microvascular HUVEC network within the printed hydrogel on day 3 (Copyright 2024 Wiely [<a href="#B255-ijms-25-12567" class="html-bibr">255</a>]).</p>
Full article ">Figure 12
<p>PEG-based photocrosslinkable bioinks. (<b>a</b>) A schematic of the structures of modified photocrosslinkable PEG derivatives for 3D bioprinting. (<b>b</b>) A schematic diagram of the process for preparing a hydrogel with cell adhesion properties. (A) Schematic representation of the ink design and the hydrogel manufacturing process. (B) Synthetic approach toward labelled RGD peptides (Copyright 2024 Wiley [<a href="#B266-ijms-25-12567" class="html-bibr">266</a>]). (<b>c</b>) A schematic illustration of the fabrication of enzymatically degradable PEG hydrogels to mimic matrix remodeling. (A) Components used in the development of the bioinspired pseudo-reversible stiffening and softening hydrogels include PEG-4-Nb (Mn∼ 5, 10, or 20 kDa), PEG-8-Nb (Mn∼40 kDa), di-thiol nondegradable linkers (PEG-2-SH; Mn∼ 2 or 3.4 kDa), di-thiol MMP degradable linker, a di-thiol MMP PEG-conjugate (PEG<sub>8</sub>MMP), and an MMP-thrombin degradable peptide linker (MMP+Thb). (B) Hydrogel tools were designed to mimic aspects of matrix degradation or matrix deposition that occurs during matrix remodeling of the cellular microenvironment through incorporation of PEG<sub>8</sub>MMP and MMP+Thb linkers, respectively. (C) A reduction of matrix density was achieved by photopolymerization of PEG hydrogels in the presence of a combination of MMP and MMP+Thb linkers, enabling triggered softening through a reduction of crosslink density upon incubation with thrombin. (D) For triggered stiffening, hydrogels were formed by photopolymerization of PEG hydrogels in the presence of MMP crosslinkers followed by secondary photopolymerization of excess reactive handles with PEG or peptide linkers (Copyright 2022 Wiley [<a href="#B267-ijms-25-12567" class="html-bibr">267</a>]). (<b>d</b>) A schematic representation of MSN bioinks for extrusion-based in situ bioprinting applications (Copyright 2023 Elsevier [<a href="#B268-ijms-25-12567" class="html-bibr">268</a>]).</p>
Full article ">Figure 13
<p>PF127 and PVA for 3D bioprinting. (<b>a</b>) A schematic diagram of the structure and gel formation mechanism of PF127 (Copyright 2020 American Chemical Society [<a href="#B283-ijms-25-12567" class="html-bibr">283</a>]). (<b>b</b>) A schematic diagram of PF127 as a sacrificial material for the preparation of microvascular tissue. (i) Schematic of the manufacturing process. (ii) Perfusion of fluorescent dextran solution into a GFP-HDFs/RFP-HUVECs co-culture construct (Copyright 2021 IOP Publishing [<a href="#B282-ijms-25-12567" class="html-bibr">282</a>]). (<b>c</b>) Norbornene-modified PVA and gelatin were used to construct a cell-laden hydrogel through volumetric bioprinting (VBP) to promote cell growth and support osteogenic differentiation. (i) (a) Schematic of the set-up for VBP. (b) Illustration of VBP of a PVA bioresin. (c) Chemical structures of norbornene-modified PVA, thiolated crosslinker (PEG2SH), and photoinitiator (LAP). (d) Mechanism of radical-mediated thiol-norbornene photoclick reaction. (ii) (a) Live(green)/dead(red) stained hMSCs following 24 h after printing, scale bars = 100 µm (i, iv). Confocal images of actin-nuclei stained hMSCs in soft and stiff gels at 24 h (ii, v) and 7 days (iii, vi) after printing, scale bars = 100 µm (ii, v) and 50 µm (iii, vi). Scale bars for all inserts are 20 µm. Visualization of single cells in soft and stiff matrix using automated IMARIS dendrite tracking, scale bars = 10µm (iii-1, vi-1). (b) Quantification of cell viability of hMSCs at different time points. (c) Quantification of average cell area in soft and stiff constructs over time. (* <span class="html-italic">p</span> = 0.0485; ns, not significant; n ≥ 3) (d) Confocal image of actin-nuclei stained hMSCs showing cell-cell contacts in the soft gels following 14 days of osteogenic culture (Copyright 2023 Wiley [<a href="#B284-ijms-25-12567" class="html-bibr">284</a>]).</p>
Full article ">Figure 14
<p>Representative photoinitiator-free photocrosslinking strategies. (<b>a</b>) A schematic diagram of UV light crosslinking based on coumarin derivatives. (<b>b</b>) A schematic diagram of UV-light-triggered imine crosslinking (Copyright 2021 American Association for the Advancement of Science [<a href="#B337-ijms-25-12567" class="html-bibr">337</a>]). (<b>c</b>) A schematic diagram of UV-light-mediated dual crosslinking based on azide-modified chitosan (Copyright 2011 American Chemical Society [<a href="#B335-ijms-25-12567" class="html-bibr">335</a>]). (<b>d</b>) A schematic diagram of photoinitiator-free photocrosslinking with SbQ as an intermediate. (i) Synthesis of PVA-SBQ and (ii) UV-light-mediated crosslinking mechanism.</p>
Full article ">Figure 15
<p>The application of NIR-light-mediated photocrosslinking based on upconversion nanoparticles (UCNPs) for in vivo 3D bioprinting. (<b>a</b>) NIR photopolymerization-based 3D printing technology that enables the noninvasive in vivo 3D bioprinting of tissue constructs (Copyright 2020 American Association for the Advancement of Science [<a href="#B340-ijms-25-12567" class="html-bibr">340</a>]). (<b>b</b>) The 3D bioprinting of noninvasive fracture scaffolds in vivo by the NIR photocuring method. (i) Schematic of the noninvasive fixation of a broken bone with the UCNPs-assisted 3D bioprinting in-vivo. (ii) Fixation scaffolds for (a) oblique and (b) comminuted fractures using UCNPs-assisted NIR 3Dprinting. Images (I, II, III, and IV) show the pre-fracture, post-fracture, 3D skeleton fixation, and corresponding magnified images of the bones respectively. The shin bones of chickens were used in the experiment, scale bar = 0.65 cm. (c) Photograph and CT image with a broken rat. (d) Bioink is subcutaneously injected into the fracture area. (e) 3D in-vivo printing. (f) Images of fracture fixation positions in-vivo, scale bar = 0.6 cm (Copyright 2024 Wiley [<a href="#B339-ijms-25-12567" class="html-bibr">339</a>]).</p>
Full article ">Scheme 1
<p>Radical generation mechanisms of type I and type II photoinitiators. * means exciting state.</p>
Full article ">
15 pages, 5676 KiB  
Article
Dynamic-Cross-Linked, Regulated, and Controllable Mineralization Degree and Morphology of Collagen Biomineralization
by Ziyao Geng, Fan Xu, Ying Liu, Aike Qiao and Tianming Du
J. Funct. Biomater. 2024, 15(12), 356; https://doi.org/10.3390/jfb15120356 - 22 Nov 2024
Abstract
The cross-linking process of collagen is one of the more important ways to improve the mineralization ability of collagen. However, the regulatory effect of dynamic cross-linking on biomineralization in vitro remains unclear. Dynamic-cross-linked mineralized collagen under different cross-linking processes, according to the process [...] Read more.
The cross-linking process of collagen is one of the more important ways to improve the mineralization ability of collagen. However, the regulatory effect of dynamic cross-linking on biomineralization in vitro remains unclear. Dynamic-cross-linked mineralized collagen under different cross-linking processes, according to the process of cross-linking and mineralization of natural bone, was prepared in this study. Mineralization was performed for 12 h at 4, 8, and 12 h of collagen cross-linking. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) showed the characteristics of dynamic-cross-linked mineralization in terms of morphological transformation and distribution. Fourier transform infrared spectroscopy (FTIR) analysis showed the crystallinity characteristics of the hydroxyapatite (HA) crystal formation. Pre-cross-linked dynamic-cross-linked mineralization refers to the process of cross-linking for a period of time and then side cross-linked mineralization. The mineral content, enzyme stability, and mechanical properties of mineralized collagen were improved through a dynamic cross-linking process of pre-cross-linking. The swelling performance was reduced through the dynamic cross-linking process of pre-cross-linking. This study suggests that the dynamic cross-linking process through pre-cross-linking could make it easier for minerals to permeate and deposit between collagen fibers, improve mineralization efficiency, and, thus, enhance the mechanical strength of biomineralization. This study can provide new ideas and a theoretical basis for designing mineralized collagen scaffolds with better bone repair ability. Full article
(This article belongs to the Special Issue Functional Composite Biomaterials for Tissue Repair)
Show Figures

Figure 1

Figure 1
<p>A schematic diagram of the collagen cross-linking and mineralization processes. (<b>a</b>) The mineralization process of natural bone is non-linear and can be divided into three stages (A) [<a href="#B13-jfb-15-00356" class="html-bibr">13</a>]. The cross-linking process of natural bone is also a nonlinear process in three stages (B). (<b>b</b>) Four sets of experiments were set up to explore the effect of dynamic cross-linking on mineralization. Group A was fully cross-linked for 12 h and mineralized for 12 h. Group B involved cross-linked mineralization that was simultaneously carried out for 12 h. Group C involved cross-linking for 4 h and then cross-linked mineralization for 12 h. Group D involved cross-linking for 8 h and then cross-linked mineralization for 12 h.</p>
Full article ">Figure 2
<p>The morphology of the pure collagen and collagen at 4 h, 8 h, 12 h, and 24 h cross-linking times. (<b>a</b>) The macroscopic morphology of the cross-linked collagen and a schematic diagram of the diffusion process of the cross-linker. The block diagram is a schematic diagram of the principles of the EDC-/NHS-cross-linked collagen. (<b>b</b>) SEM images of the cross-linked collagen at the 1 mm (×100), 200 μm (×500), and 50 μm (×1500) scales. The red dashed boxes represent areas of the image that have increased magnification. (<b>A1</b>–<b>D1</b>) show collagen cross-linking diagrams. The green curve indicates the collagen fibers, and the black dashed line indicates the degree of cross-linking. The degree of cross-linking increased with increases in time. (<b>A1</b>–<b>A4</b>) show the cross-linking collagen for 4 h. (<b>B1</b>–<b>B4</b>) show the cross-linking collagen for 8 h. (<b>C1</b>–<b>C4</b>) show the cross-linking collagen for 12 h. The fibers had a smooth surface (<b>C4</b>, circle). (<b>D1</b>–<b>D4</b>) shows the cross-linking collagen for 24 h.</p>
Full article ">Figure 3
<p>(<b>a</b>) The FTIR of collagen at 4 h, 8 h, 12 h, and 24 h cross-linking times. (<b>b</b>) The enzymolysis rate of collagen at 4 h, 8 h, 12 h, and 24 h cross-linking times. (<b>c</b>) The Young’s modulus of collagen at 4 h, 8 h, 12 h, and 24 h cross-linking times. ** <span class="html-italic">p</span> &lt; 0.01, ns: no significant difference.</p>
Full article ">Figure 4
<p>The morphologies of the mineralized collagen under different cross-linking conditions. (<b>a</b>) SEM images of the mineralized collagen at 1 mm, 200 μm, and 50 μm scales. A1–A3 are the images of Group A. B1–B3 are the images of Group B. C1–C3 are the images of Group C. D1–D3 are the images of Group D. Hydroxyapatite was found in all four groups. (A3–D3, triangles). (<b>b</b>) A schematic diagram of the mineralization process under different cross-linking conditions. Group A is mineralization after complete cross-linking. Group B is simultaneous cross-linked mineralization. Group C&amp;D is simultaneous cross-linked mineralization after pre-cross-linking. (<b>c</b>) EDS images of mineralized collagen at 1 mm. C stands for carbon, Ca stands for calcium, and P stands for phosphorus. A4–A6 are the images of Group A. B4–B6 are the images of Group B. C4–C6 are the images of Group C. D4–D6 are the images of Group D. (<b>d</b>) A schematic diagram showing that the mineral mass of the dynamic cross-linking group was more deeply embedded in the mineralized collagen, while the mineral mass of the mineralized after cross-linking group was less and was more attached to the surface.</p>
Full article ">Figure 5
<p>(<b>a</b>) The thermogravimetric curves of the mineralized collagen matrix. The yellow box I indicates a range of 100–180 °C, and the green box II indicates a range of 200–550 °C. (<b>b</b>) The mass fraction of each component of the composite mineralized collagen. (<b>c</b>) The enzymolysis rate of the mineralized collagen. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 6
<p>(<b>a</b>) TEM and SAED images of the cross-linked and then mineralized collagen and dynamic cross-linking of the mineralized collagen. (<b>A</b>–<b>C</b>) are the group of “Mineralization occurs after complete crosslinking”. (<b>D</b>–<b>F</b>) are the group of “Cross-linked mineralization occurs simultaneously”. (<b>b</b>) The FTIR of the mineralized collagen in each group. (<b>c</b>) The amplification of the gray area of (<b>b</b>) (500 cm<sup>−1</sup>–1500 cm<sup>−1</sup>).</p>
Full article ">Figure 7
<p>(<b>a</b>) Comparison of the mineralized collagen before and after swelling under different cross-linking conditions. (<b>A1</b>–<b>D1</b>) are the images before swelling, and (<b>A2</b>–<b>D2</b>) are the images after swelling. (<b>A1</b>–<b>A2</b>) are the images of Group A. (<b>B1</b>–<b>B2</b>) are the images of Group B. (<b>C1</b>–<b>C2</b>) are the images of Group C. (<b>D1</b>–<b>D2</b>) are the images of Group D. (<b>b</b>) The swelling rate of the mineralized collagen. (<b>c</b>) The Young’s modulus of the mineralized collagen. (<b>d</b>) The Young’s modulus of the mineralized collagen after swelling. (<b>e</b>) A schematic diagram of the dynamic-cross-linked mineralization after pre-cross-linking. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, ns: no significant difference.</p>
Full article ">
20 pages, 12267 KiB  
Article
Biocompatibility Analysis of the Silver-Coated Microporous Titanium Implants Manufactured with 3D-Printing Technology
by Maxim Shevtsov, Emil Pitkin, Stephanie E. Combs, Natalia Yudintceva, Denis Nazarov, Greg Van Der Meulen, Chris Preucil, Michael Akkaoui and Mark Pitkin
Nanomaterials 2024, 14(23), 1876; https://doi.org/10.3390/nano14231876 - 22 Nov 2024
Abstract
3D-printed microporous titanium scaffolds enjoy good biointegration with the residuum’s soft and bone tissues, and they promote excellent biomechanical properties in attached prostheses. Implant-associated infection, however, remains a major clinical challenge. Silver-based implant coatings can potentially reduce bacterial growth and inhibit biofilm formation, [...] Read more.
3D-printed microporous titanium scaffolds enjoy good biointegration with the residuum’s soft and bone tissues, and they promote excellent biomechanical properties in attached prostheses. Implant-associated infection, however, remains a major clinical challenge. Silver-based implant coatings can potentially reduce bacterial growth and inhibit biofilm formation, thereby reducing the risk of periprosthetic infections. In the current study, a 1-µm thick silver coating was prepared on the surface of a 3D-printed microporous titanium alloy with physical vapor deposition (PVD), with a final silver content of 1.00 ± 02 mg/cm2. Cell viability was evaluated with an MTT assay of MC3T3-E1 osteoblasts and human dermal fibroblasts cultured on the surface of the implants, and showed low cytotoxicity for cells during the 14-day follow-up period. Quantitative real-time polymerase chain reaction (RT-PCR) analysis of the relative gene expression of the extracellular matrix components (fibronectin, vitronectin, type I collagen) and cell adhesion markers (α2, α5, αV, β1 integrins) in dermal fibroblasts showed that cell adhesion was not reduced by the silver coating of the microporous implants. An RT-PCR analysis of gene expression related to osteogenic differentiation, including TGF-β1, SMAD4, osteocalcin, osteopontin, and osteonectin in MC3T3-E1 osteoblasts, demonstrated that silver coating did not reduce the osteogenic activity of cells and, to the contrary, enhanced the activity of the TGF-β signaling pathway. For representative sample S5 on day 14, the gene expression levels were 7.15 ± 0.29 (osteonectin), 6.08 ± 0.12 (osteocalcin), and 11.19 ± 0.77 (osteopontin). In conclusion, the data indicate that the silver coating of the microporous titanium implants did not reduce the biointegrative or osteoinductive properties of the titanium scaffold, a finding that argues in favor of applying this coating in designing personalized osseointegrated implants. Full article
Show Figures

Figure 1

Figure 1
<p>Tablets for the study: (<b>A</b>) set of tablets (S1–S9) fabricated with 3D-printing technology and coated with silver; (<b>B</b>) r<sub>1</sub> is the outer radius of the tablets and r<sub>2</sub> is the radius of a central solid core.</p>
Full article ">Figure 2
<p>MTT assay of dermal fibroblasts and MC3T3-E1 osteoblast cells on silver-coated 3D-printed titanium microporous implants (S1–S9). Cell viability (%) was evaluated on the 1st, 3rd, 7th, and 14th day after co-incubation. Sintered Ti implant and 3D-printed implant without silver coating were used as controls. Data is presented from three independent experiments as M ± SD.</p>
Full article ">Figure 3
<p>Representative scanning electron microscopy images of MC3T3-E1 cells and fibroblasts cultured on the samples S5 following 72 h of co-incubation.</p>
Full article ">Figure 4
<p>Comparison of expression of integrins and extracellular matrix component (fibronectin, vitronectin, type I collagen) genes of dermal fibroblasts on silver-coated 3D-printed titanium implants S1–S9 4, 24, 48, and 72 h after co-culturing. Analysis of gene expression related to fibronectin, vitronectin, and type I collagen was performed following 4 and 72 h of co-culturing cells on the surface of implants. Data is presented from three independent experiments as M ± SD. <span class="html-italic">p</span> &lt; 0.01 for testing mean expression levels.</p>
Full article ">Figure 5
<p>Comparison of gene (FAK, vinculin, paxillin) expression for MC3T3-E1 cells co-cultured on silver-coated titanium implants with various pore sizes (S1–S9) after 1, 3, 7, and 14 days. Data is presented from three independent experiments as M ± SD. <span class="html-italic">p</span> &lt; 0.01 for testing mean expression levels.</p>
Full article ">Figure 6
<p>Comparison of expression of osteocalcin, osteopontin, and osteocalcin genes of MC3T3-E1 cells following co-incubation with silver-coated titanium implants (S1–S9) after 1, 3, 7, and 14 days. Analysis of TGF-β1 and SMAD4 gene expression in MC3T3-E1 osteoblast cells was performed on days 1 and 7 after co-incubation. Data is presented from three independent experiments as M ± SD. <span class="html-italic">p</span> &lt; 0.01 for testing mean expression levels.</p>
Full article ">
18 pages, 3119 KiB  
Review
Synthesis of Diazacyclic and Triazacyclic Small-Molecule Libraries Using Vicinal Chiral Diamines Generated from Modified Short Peptides and Their Application for Drug Discovery
by Mukund P. Tantak, Ramanjaneyulu Rayala, Prakash Chaudhari, Chhanda C. Danta and Adel Nefzi
Pharmaceuticals 2024, 17(12), 1566; https://doi.org/10.3390/ph17121566 - 22 Nov 2024
Viewed by 100
Abstract
Small-molecule probes are powerful tools for studying biological systems and can serve as lead compounds for developing new therapeutics. Especially, nitrogen heterocycles are of considerable importance in the pharmaceutical field. These compounds are found in numerous bioactive structures. Their synthesis often requires several [...] Read more.
Small-molecule probes are powerful tools for studying biological systems and can serve as lead compounds for developing new therapeutics. Especially, nitrogen heterocycles are of considerable importance in the pharmaceutical field. These compounds are found in numerous bioactive structures. Their synthesis often requires several steps or the use of functionalized starting materials. This review describes the use of vicinal diamines generated from modified short peptides to access substituted diaza- and triazacyclic compounds. Small-molecule diaza- and triazacyclic compounds with different substitution patterns and embedded in various molecular frameworks constitute important structure classes in the search for bioactivity. The compounds are designed to follow known drug likeness rules, including “Lipinski’s Rule of Five”. The screening of diazacyclic and traizacyclic libraries has shown the utility of these classes of compounds for the de novo identification of highly active compounds, including antimalarials, antimicrobial compounds, antifibrotic compounds, potent analgesics, and antitumor agents. Examples of the synthesis of diazacyclic and triazacyclic small-molecule libraries from vicinal chiral polyamines generated from modified short peptides and their application for the identification of highly active compounds are described. Full article
(This article belongs to the Special Issue Nitrogen Containing Scaffolds in Medicinal Chemistry 2023)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Deconvolution strategy of the positional scanning library of 1,5-disubstituted acylated 2-amino-4,5-dihydroimidazoles for the identification of RORγ inhibitors.</p>
Full article ">Figure 2
<p>Examples of identified diazacyclic active compounds.</p>
Full article ">Figure 3
<p>Examples of identified active bicyclic guanidines.</p>
Full article ">Figure 4
<p>Active diimidazodiazepine compounds and their affinities for μ (MOR), δ (DOR), and κ (KOR) opioid receptors.</p>
Full article ">Figure 5
<p>Examples of identified bis diazacyclic active compounds.</p>
Full article ">Scheme 1
<p>Parallel synthesis of trisubstituted diazacyclic libraries from resin-bound vicinal secondary diamines.</p>
Full article ">Scheme 2
<p>Synthesis of 1,5-disubstituted acylated 2-amino-4,5-dihydroimidazoles from resin-bound vicinal amines.</p>
Full article ">Scheme 3
<p>Synthesis of monoketopiperazines from secondary vicinal diamines.</p>
Full article ">Scheme 4
<p>Parallel synthesis of bicyclic guanidine from resin-bound vicinal triamines.</p>
Full article ">Scheme 5
<p>Synthesis of diaza-6-azoniaspiro compounds from resin-bound vicinal triamines.</p>
Full article ">Scheme 6
<p>Synthesis of triazepin-amines from resin-bound vicinal triamines.</p>
Full article ">Scheme 7
<p>Combined solid-phase and solution-phase synthesis of fused diimidazodiazepines from vicinal tetraamines.</p>
Full article ">Scheme 8
<p>Solid-phase synthesis of pyrrolidine bis diazacyclic libraries from pyrrolidine di-vicinal diamines.</p>
Full article ">Scheme 9
<p>Solid-phase synthesis of bis diazacyclic libraries from di-vicinal diamines.</p>
Full article ">Scheme 10
<p>Solid-phase synthesis of imidazoline-tethered diazacyclic libraries from di-vicinal diamines.</p>
Full article ">Scheme 11
<p>Solid-phase synthesis of bis cyclic diazacyclic libraries from vicinal tetraamines.</p>
Full article ">Scheme 12
<p>Synthesis of oligodiazacyclic compounds from oligo vicinal diamines.</p>
Full article ">Scheme 13
<p>Synthesis of oligodiazacyclic compounds from oligo vicinal diamines.</p>
Full article ">Scheme 14
<p>Synthesis of macrocyclic-containing diazacyclic compounds from resin-bound oligo vicinal diamines.</p>
Full article ">
20 pages, 4201 KiB  
Article
Impact of Particle Size and Sintering Temperature on Calcium Phosphate Gyroid Structure Scaffolds for Bone Tissue Engineering
by Romina Haydeé Aspera-Werz, Guanqiao Chen, Lea Schilonka, Islam Bouakaz, Catherine Bronne, Elisabeth Cobraiville, Grégory Nolens and Andreas Nussler
J. Funct. Biomater. 2024, 15(12), 355; https://doi.org/10.3390/jfb15120355 - 21 Nov 2024
Viewed by 261
Abstract
Due to the chemical composition and structure of the target tissue, autologous bone grafting remains the gold standard for orthopedic applications worldwide. However, ongoing advancements in alternative grafting materials show that 3D-printed synthetic biomaterials offer many advantages. For instance, they provide high availability, [...] Read more.
Due to the chemical composition and structure of the target tissue, autologous bone grafting remains the gold standard for orthopedic applications worldwide. However, ongoing advancements in alternative grafting materials show that 3D-printed synthetic biomaterials offer many advantages. For instance, they provide high availability, have low clinical limitations, and can be designed with a chemical composition and structure comparable to the target tissue. This study aimed to compare the influences of particle size and sintering temperature on the mechanical properties and biocompatibility of calcium phosphate (CaP) gyroid scaffolds. CaP gyroid scaffolds were fabricated by 3D printing using powders with the same chemical composition but different particle sizes and sintering temperatures. The physicochemical characterization of the scaffolds was performed using X-ray diffractometry, scanning electron microscopy, and microtomography analyses. The immortalized human mesenchymal stem cell line SCP-1 (osteoblast-like cells) and osteoclast-like cells (THP-1 cells) were seeded on the scaffolds as mono- or co-cultures. Bone cell attachment, number of live cells, and functionality were assessed at different time points over a period of 21 days. Improvements in mechanical properties were observed for scaffolds fabricated with narrow-particle-size-distribution powder. The physicochemical analysis showed that the microstructure varied with sintering temperature and that narrow particle size distribution resulted in smaller micropores and a smoother surface. Viable osteoblast- and osteoclast-like cells were observed for all scaffolds tested, but scaffolds produced with a smaller particle size distribution showed less attachment of osteoblast-like cells. Interestingly, low attachment of osteoclast-like cells was observed for all scaffolds regardless of surface roughness. Although bone cell adhesion was lower in scaffolds made with powder containing smaller particle sizes, the long-term function of osteoblast-like and osteoclast-like cells was superior in scaffolds with improved mechanical properties. Full article
Show Figures

Figure 1

Figure 1
<p>Scaffold with cylindrical shape and gyroid structure. (<b>a</b>) Top view. (<b>b</b>) Side view.</p>
Full article ">Figure 2
<p>Powder composition determined by X-ray diffractometry (XRD). Representative XRD curve for (<b>a</b>) powder A and (<b>b</b>) powder B.</p>
Full article ">Figure 3
<p>Mechanical characterization of the three scaffolds tested. (<b>a</b>) Maximum force, (<b>b</b>) maximum stress, (<b>c</b>) displacements at maximum load, and (<b>d</b>) flexural strength were analyzed on scaffolds generated with powder A sintering at 1230 °C [Scaffold A] or powder B sintering at 1250 °C and 1210 °C [scaffolds B<sub>I</sub> and B<sub>II</sub>, respectively]. The data are presented as the mean, standard error of the mean, and all data points. Data were analyzed by the Kruskal–Wallis test followed by Dunn’s multiple comparisons. <span class="html-italic">p</span>-values are classified as * <span class="html-italic">p</span> &lt; 0.05; *** <span class="html-italic">p</span> &lt; 0.001; and **** <span class="html-italic">p</span> &lt; 0.0001 for comparison between scaffold B and scaffold A and as # <span class="html-italic">p</span> &lt; 0.05 for comparison within scaffold B.</p>
Full article ">Figure 4
<p>Surface topographies of the three scaffolds analyzed by scanning electron microscopy (SEM). (<b>a</b>) Scaffold generated with powder A and sintering at 1230 °C. (<b>b</b>) Scaffold generated with powder B and sintering at 1250 °C. (<b>c</b>) Scaffold generated with powder B and sintering at 1210 °C (scale bar 10 µm).</p>
Full article ">Figure 5
<p>SCP-1 cell attachment, number of live cells, and proliferation on three scaffolds tested. SCP-1 cells were seeded and cultured on scaffolds A, B<sub>I</sub>, and B<sub>II</sub> for 21 days. (<b>a</b>) Attached SCP-1 cells on scaffolds compared to cultured polystyrene. Number of live SCP-1 cells were analyzed after 24 h, 48 h, 7 days, 14 days, and 21 days by total DNA levels (<b>b</b>) and visualized by esterase activity (<b>c</b>) using calcein-AM (green) and nuclear staining using Hoechst 33342 (blue) (scale bar 2000 µm). Each measure was conducted at least three independent times in triplicate. The data are presented as the mean, standard error of the mean, and all data points. Data were analyzed by the Kruskal–Wallis test followed by Dunn’s multiple comparisons (<b>a</b>) or a two-way analysis of variance test followed by Tukey’s multiple comparisons (<b>b</b>). <span class="html-italic">p</span>-values are classified as ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001; and **** <span class="html-italic">p</span> &lt; 0.0001 for comparison between scaffold B and scaffold A and as ## <span class="html-italic">p</span> &lt; 0.01; ### <span class="html-italic">p</span> &lt; 0.001; and #### <span class="html-italic">p</span> &lt; 0.0001 for comparison within scaffold B.</p>
Full article ">Figure 6
<p>SCP-1 osteogenic differentiation potential on three scaffolds tested. SCP-1 cells were seeded and cultured under osteogenic condition on scaffolds A, BI, and BII for 21 days. (<b>a</b>) Metabolic activity of SCP-1 cells were analyzed after 24 h, 48 h, 7 days, 14 days, and 21 days by mitochondrial activity as relative fluorescence units (RFU). (<b>b</b>) Alkaline phosphatase (AP) activity normalized to DNA of SCP-1 cells were analyzed after 7 days, 14 days, and 21 days as relative absorbance units (RAU). (<b>c</b>) Procollagen type I N-propeptide (PINP) supernatant levels were determined after 21-day osteogenic culture. Each measure was conducted at least three independent times in duplicate. The data are presented as the mean, standard error of the mean, and all data points. Data were analyzed by a two-way analysis of variance test followed by Tukey’s multiple comparisons. <span class="html-italic">p</span>-values are classified as * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; and **** <span class="html-italic">p</span> &lt; 0.0001 for comparison between scaffold B and scaffold A and as # <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; and #### <span class="html-italic">p</span> &lt; 0.0001 for comparison within scaffold B.</p>
Full article ">Figure 7
<p>THP-1 cell attachment and number of live cells on three scaffolds tested. THP-1 cells were seeded and cultured on scaffolds A, BI, and BII for 24 h. (<b>a</b>) Attached SCP-1 cells on scaffolds compared to cultured polystyrene. Number of live THP-1 cells were visualized after 24 h by esterase activity (<b>b</b>) using calcein-AM (green) and nuclear staining using Hoechst 33342 (blue) (scale bar 2000 µm). Each measure was conducted at least three independent times in triplicate. The data are presented as the mean, standard error of the mean, and all data points. Data were analyzed by the Kruskal–Wallis test followed by Dunn’s multiple comparisons. <span class="html-italic">p</span>-values are classified as **** <span class="html-italic">p</span> &lt; 0.0001 for comparison between scaffold B and scaffold A.</p>
Full article ">Figure 8
<p>Bone cell viability and proliferation in co-cultures on the three scaffolds tested. THP-1 and SCP-1 were seeded and co-cultured on scaffolds A, BI, and BII for 21 days. Quantification of total DNA (<b>a</b>) and mitochondrial activity by resazurin conversion (<b>b</b>) in bone co-cultures after 7 days, 14 days, and 21 days. Number of live bone co-cultures were visualized by esterase activity (<b>c</b>) using calcein-AM (green) and nuclear staining with Hoechst 33342 (blue) (scale bar 2000 µm). Each measure was conducted at least three independent times in triplicate. The data are presented as the mean, standard error of the mean, and all data points. Data were analyzed by two-way analysis of variance test followed by Tukey’s multiple comparisons. <span class="html-italic">p</span>-values are classified as **** <span class="html-italic">p</span> &lt; 0.0001 for comparison between scaffold B and scaffold A and as # <span class="html-italic">p</span> &lt; 0.05 for comparison within scaffold B.</p>
Full article ">Figure 9
<p>Osteoblast- and osteoclast-like cell function in co-cultures on the three scaffolds tested. THP-1 and SCP-1 were seeded and co-cultured on scaffolds A, BI, and BII for 21 days. (<b>a</b>) Alkaline phosphatase (AP), (<b>b</b>) carbonic anhydrase II (CAII), and (<b>c</b>) tartrate-resistant acid phosphatase (TRAP) activity normalized to DNA of bone co-cultures were analyzed after 7 days, 14 days, and 21 days as relative absorbance units (RAU). (<b>d</b>) Procollagen type I N-propeptide (PINP) and collagen type I N-telopeptide (NTX) supernatant levels were determined after a 21-day culture. Each measure was conducted at least three independent times in duplicate. The data are presented as the mean or standard error of the mean. Data were analyzed by a two-way analysis of variance test followed by Tukey’s multiple comparisons. <span class="html-italic">p</span>-values are classified as **** <span class="html-italic">p</span> &lt; 0.0001, ** <span class="html-italic">p</span> &lt; 0.01, and * <span class="html-italic">p</span> &lt; 0.05 for comparison between scaffold B and scaffold A and as ## <span class="html-italic">p</span> &lt; 0.01 for comparison within scaffold B.</p>
Full article ">
24 pages, 7292 KiB  
Article
The Impact of Temperature and Pressure on the Structural Stability of Solvated Solid-State Conformations of Bombyx mori Silk Fibroins: Insights from Molecular Dynamics Simulations
by Ezekiel Edward Nettey-Oppong, Riaz Muhammad, Ahmed Ali, Hyun-Woo Jeong, Young-Seek Seok, Seong-Wan Kim and Seung Ho Choi
Materials 2024, 17(23), 5686; https://doi.org/10.3390/ma17235686 - 21 Nov 2024
Viewed by 357
Abstract
Bombyx mori silk fibroin is a promising biopolymer with notable mechanical strength, biocompatibility, and potential for diverse biomedical applications, such as tissue engineering scaffolds, and drug delivery. These properties are intrinsically linked to the structural characteristics of silk fibroin, making it essential to [...] Read more.
Bombyx mori silk fibroin is a promising biopolymer with notable mechanical strength, biocompatibility, and potential for diverse biomedical applications, such as tissue engineering scaffolds, and drug delivery. These properties are intrinsically linked to the structural characteristics of silk fibroin, making it essential to understand its molecular stability under varying environmental conditions. This study employed molecular dynamics simulations to examine the structural stability of silk I and silk II conformations of silk fibroin under changes in temperature (298 K to 378 K) and pressure (0.1 MPa to 700 MPa). Key parameters, including Root Mean Square Deviation (RMSD), Root Mean Square Fluctuation (RMSF), and Radius of Gyration (Rg) were analyzed, along with non-bonded interactions such as van der Waals and electrostatic potential energy. Our findings demonstrate that both temperature and pressure exert a destabilizing effect on silk fibroin, with silk I exhibiting a higher susceptibility to destabilization compared to silk II. Additionally, pressure elevated the van der Waals energy in silk I, while temperature led to a reduction. In contrast, electrostatic potential energy remained unaffected by these environmental conditions, highlighting stable long-range interactions throughout the study. Silk II’s tightly packed β-sheet structure offers greater resilience to environmental changes, while the more flexible α-helices in silk I make it more susceptible to structural perturbations. These findings provide valuable insights into the atomic-level behavior of silk fibroin, contributing to a deeper understanding of its potential for applications in environments where mechanical or thermal stress is a factor. The study underscores the importance of computational approaches in exploring protein stability and supports the continued development of silk fibroin for biomedical and engineering applications. Full article
(This article belongs to the Special Issue Advances in Bio-Polymer and Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of <span class="html-italic">Bombyx mori</span> silk structure. During the pupa stage of their metamorphosis into moths, silkworms spin silk fibers to construct protective cocoons. Each silk fiber is composed of two core fibroin filaments encased by sericin (depicted in purple), an adhesive glycoprotein that facilitates fiber cohesion. At the molecular level, each fibroin filament consists of numerous assemblies of nanofibrils, which can adopt either silk I or silk II structural conformations. These structural forms are determined by the specific arrangement of secondary protein structures within the fibroin. The silk I structure predominantly features type II β-turns and α-helices, while the silk II structure is mainly characterized by β-turns and β-sheets. Both of these secondary structures contribute to the fiber’s mechanical properties and functional versatility across various applications.</p>
Full article ">Figure 2
<p>Schematic of the secondary structures of <span class="html-italic">Bombyx mori</span> silk fibroin. The silk fibroin protein comprises several distinct secondary structures that play a critical role in determining the material’s mechanical strength and biological properties. These secondary structures include the following: (<b>a</b>) α-helix, a right-handed coiled structure that contributes to flexibility; (<b>b</b>) β-sheet, an extended conformation that forms the crystalline regions, providing mechanical robustness; and (<b>c</b>) random coils, which are unstructured regions that contribute to the amorphous domains of the protein. The schematic also illustrates the specific amino acid residues associated with each of these secondary structures, alongside the corresponding all-atom models, providing a molecular-level perspective on silk fibroin’s hierarchical organization.</p>
Full article ">Figure 3
<p>Schematic representation of the primary structure of <span class="html-italic">Bombyx mori</span> silk fibroin. The heavy chain of silk fibroin consists of alternating crystalline (R1 to R12) and amorphous phases (L1 to L11), along with N terminus (N) and C terminus (C), each having distinct structural and functional roles. The crystalline regions, indicated by the R domains, are primarily composed of β-sheet structures that impart rigidity and strength to the fiber. In contrast, the amorphous regions, denoted as L domains, serve as flexible linkers that provide elasticity and contribute to the material’s overall mechanical performance. This schematic highlights the representative molecular structure used for simulations, where domain R6 (crystalline) is flanked by domains L5 and L6 (amorphous), providing a balanced model for studying both phases of the fibroin.</p>
Full article ">Figure 4
<p>Molecular dynamics equilibration simulation of silk fibroin protein structures. This figure illustrates the evolution of both potential energy and kinetic energy over time during the equilibration phase for hydrated silk fibroin systems. Equilibration was performed under the NPT ensemble for the silk I (<b>a</b>) and silk II (<b>b</b>) structures. The gradual decrease and stabilization of the total potential energy throughout the simulation indicates that the system achieves a stable configuration as the atoms settle into their equilibrium positions. The kinetic energy, in contrast, remains relatively constant, reflecting consistent thermal motion within the system. This energy evolution demonstrates the successful stabilization of silk fibroin structures under the specified conditions, preparing them for further simulation analyses.</p>
Full article ">Figure 5
<p>Volume changes of the simulation cells during equilibration of silk fibroin structures. The figure presents the observed reduction in the volume of the simulation cells during equilibration. For the silk I structure (<b>a</b>), the cubic cell size decreased from an initial length of 105 Å to 102.7 Å, while for the silk II structure (<b>b</b>), the cell size reduced from 89 Å to 86.9 Å. This volume contraction is indicative of the system reaching equilibrium, as the protein and water molecules reorganize into a more compact and energetically favorable configuration. The simulations ensured complete hydration of the protein by maintaining a 10 Å buffer between the protein structures and the cell boundaries. A visual snapshot on the far right provides a cross-sectional view of the hydrated silk I and silk II proteins, illustrating the distribution of water molecules around the protein structures and confirming that the proteins are fully solvated within the simulation cells.</p>
Full article ">Figure 6
<p>Root Mean Square Deviation (RMSD) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) RMSD values for the backbone atoms of silk I across a pressure range of 0.1 MPa to 700 MPa, showing how the structural deviation increases with pressure. (<b>b</b>) RMSD values for the backbone atoms of silk II under the same pressure range. Both structures exhibit increasing deviation with pressure. (<b>c</b>) The average RMSD values for silk I and silk II as a function of pressure. Silk I has a minimum deviation of 0.804 Å at 0.1 MPa, increasing to a maximum of 1.454 Å at 700 MPa. Similarly, silk II exhibits a minimum deviation of 0.772 Å at 0.1 MPa and a maximum of 1.285 Å at 700 MPa. These results demonstrate that pressure induces structural instability, with silk I experiencing a more pronounced deviation than silk II. (<b>d</b>) RMSD of silk I backbone atoms across a temperature range of 298 K to 378 K, indicating how thermal agitation affects structural deviation. (<b>e</b>) RMSD values for silk II at the same temperature range, illustrating the temperature-dependent structural perturbations. (<b>f</b>) The average RMSD values for silk I and silk II as a function of temperature. The temperature increase caused a rise in RMSD for both structures, with silk I showing a deviation from 0.764 Å at 298 K to 0.871 Å at 378 K, and silk II deviating from 0.742 Å to 0.843 Å over the same temperature range. This increase indicates that thermal agitation leads to greater atomic movement and structural deviations, particularly for silk I.</p>
Full article ">Figure 7
<p>Root Mean Square Fluctuation (RMSF) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) RMSF values for silk I backbone atoms across a pressure range of 0.1 MPa to 700 MPa, showing per-residue fluctuations and the effect of pressure on protein flexibility. (<b>b</b>) RMSF values for silk II under the same pressure range, highlighting differences in flexibility between the two structures. (<b>c</b>) The average RMSF values for silk I and silk II as a function of pressure. The minimum fluctuation for silk I was 0.470 Å at 0.1 MPa, increasing to a maximum of 0.507 Å at 700 MPa. For silk II, the fluctuation values ranged from 0.454 Å to 0.453 Å over the same pressure range. The observed fluctuations are relatively low, indicating minimal atomic mobility under pressure for both structures. (<b>d</b>) RMSF values for silk I backbone atoms across a temperature range of 298 K to 378 K, showing increased fluctuations with rising temperature. (<b>e</b>) RMSF values for silk II at the same temperature range, reflecting a similar trend of increased fluctuations with temperature. (<b>f</b>) The average RMSF values for silk I and silk II as a function of temperature. The fluctuations increase with temperature, with silk I showing a rise from 0.498 Å at 298 K to 0.592 Å at 378 K, and silk II increasing from 0.468 Å to 0.536 Å. These results suggest that temperature-induced thermal agitation leads to increased per-residue flexibility, particularly for silk I, which demonstrates greater fluctuation than silk II.</p>
Full article ">Figure 8
<p>Radius of Gyration (R<sub>g</sub>) of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) R<sub>g</sub> values for silk I across a pressure range of 0.1 MPa to 700 MPa, illustrating how pressure affects the overall compactness of the structure. (<b>b</b>) R<sub>g</sub> values for silk II under the same pressure range, showing a similar trend but with lower values compared to silk I. (<b>c</b>) The average R<sub>g</sub> values for silk I and silk II as a function of pressure. An increase in pressure leads to a reduction in R<sub>g</sub>, indicating increased compaction of both structures. Silk I shows a maximum R<sub>g</sub> of 22.485 Å at 0.1 MPa, decreasing to 21.520 Å at 700 MPa. Silk II exhibits a maximum R<sub>g</sub> of 20.635 Å at 0.1 MPa, reducing to 19.929 Å at 700 MPa. These results confirm that pressure induces compaction, with silk I showing a greater reduction in compactness than silk II. (<b>d</b>) R<sub>g</sub> values for silk I across a temperature range of 298 K to 378 K, indicating how thermal effects impact protein packing. (<b>e</b>) R<sub>g</sub> values for silk II under the same temperature range, showing minimal changes in compactness. (<b>f</b>) The average R<sub>g</sub> values for silk I and silk II as a function of temperature. Both structures show minimal alterations in compactness with temperature, with R<sub>g</sub> values of 22.762 Å for silk I and 20.874 Å for silk II at 298 K, slightly decreasing to 22.711 Å and 20.937 Å, respectively, at 378 K. The small changes in R<sub>g</sub> suggest that the temperature range studied has a negligible effect on the compactness of the silk structures.</p>
Full article ">Figure 9
<p>Non-bonded interactions of silk I and silk II structures under different pressure and temperature conditions. (<b>a</b>) The average van der Waals energy as a function of pressure, from 0.1 MPa to 700 MPa. For both silk I and silk II, the van der Waals energy increased with rising pressure, reflecting the compression of inter-atomic distances and the strengthening of long-range non-bonded interactions. (<b>b</b>) The average electrostatic potential energy as a function of pressure. The electrostatic potential energy remained constant for both silk I and silk II, with values of −1.97 × 10<sup>6</sup> kcal/mol and −1.16 × 10<sup>6</sup> kcal/mol, respectively, indicating that pressure has no significant effect on electrostatic interactions. (<b>c</b>) The average van der Waals energy as a function of temperature, from 298 K to 378 K. In contrast to pressure, the van der Waals energy decreased with increasing temperature for both silk structures, due to the expansion of inter-atomic distances and the weakening of non-bonded interactions. (<b>d</b>) The average electrostatic potential energy as a function of temperature. Similar to pressure, the electrostatic potential energy remained constant with temperature for both silk I and silk II, highlighting the stability of long-range electrostatic interactions under thermal fluctuations.</p>
Full article ">
19 pages, 2233 KiB  
Article
Structure–Activity Relationship Studies in a Series of 2-Aryloxy-N-(pyrimidin-5-yl)acetamide Inhibitors of SLACK Potassium Channels
by Nigam M. Mishra, Brittany D. Spitznagel, Yu Du, Yasmeen K. Mohamed, Ying Qin, C. David Weaver and Kyle A. Emmitte
Molecules 2024, 29(23), 5494; https://doi.org/10.3390/molecules29235494 - 21 Nov 2024
Viewed by 280
Abstract
Epilepsy of infancy with migrating focal seizures (EIMFS) is a rare, serious, and pharmacoresistant epileptic disorder often linked to gain-of-function mutations in the KCNT1 gene. KCNT1 encodes the sodium-activated potassium channel known as SLACK, making small molecule inhibitors of SLACK channels a compelling [...] Read more.
Epilepsy of infancy with migrating focal seizures (EIMFS) is a rare, serious, and pharmacoresistant epileptic disorder often linked to gain-of-function mutations in the KCNT1 gene. KCNT1 encodes the sodium-activated potassium channel known as SLACK, making small molecule inhibitors of SLACK channels a compelling approach to the treatment of EIMFS and other epilepsies associated with KCNT1 mutations. In this manuscript, we describe a hit optimization effort executed within a series of 2-aryloxy-N-(pyrimidin-5-yl)acetamides that were identified via a high-throughput screen. We systematically prepared analogs in four distinct regions of the scaffold and evaluated their functional activity in a whole-cell, automated patch clamp (APC) assay to establish structure-activity relationships for wild-type (WT) SLACK inhibition. Two selected analogs were also profiled for selectivity versus other members of the Slo family of potassium channels, of which SLACK is a member, and versus a panel of structurally diverse ion channels. The same two analogs were evaluated for activity versus the WT mouse channel as well as two clinically relevant mutant human channels. Full article
Show Figures

Figure 1

Figure 1
<p>Examples of small molecule SLACK inhibitors that have been reported in the literature. Reported IC<sub>50</sub> values obtained using whole-cell, patch clamp electrophysiology versus WT SLACK. Refer to original references cited in the text for experimental details, as cell lines and protocols differ.</p>
Full article ">Figure 2
<p>Hit optimization plan for SLACK inhibitor <b>10</b> (VU0545326) in four regions.</p>
Full article ">Scheme 1
<p>Synthesis of analogs. <span class="html-italic">Reagents and Conditions:</span> (<b>a</b>) ArXH, K<sub>2</sub>CO<sub>3</sub>, CH<sub>3</sub>CN, µwave, 80 °C, 30 min (For X=NMe, an additional reaction with K<sub>2</sub>CO<sub>3</sub>, MeI, DMF was employed); (<b>b</b>) H<sub>2</sub>, Pt(sulfided)/C, MeOH, 70 °C, H-Cube<sup>®</sup>; (<b>c</b>) (4-chlorophenyl)-ZYCO<sub>2</sub>H, HATU, DIEA, DMF; (<b>d</b>) 2-bromo-2-methylpropionyl bromide, DIEA, CH<sub>2</sub>Cl<sub>2</sub>; (<b>e</b>) ArZH, CuBr·SMe<sub>2</sub>, PCy<sub>3</sub>, K<sub>3</sub>PO<sub>4</sub>, CH<sub>3</sub>CN.</p>
Full article ">Scheme 2
<p>Synthesis of analogs. Reagents and Conditions: (<b>a</b>) HN(<span class="html-italic">i</span>-Pr)<sub>2</sub>, <span class="html-italic">n</span>-BuLi (1.6M in hexanes), THF, −78 °C, then 4-chlorobenzyl bromide, 55%; (<b>b</b>) NaOH, MeOH, H<sub>2</sub>O, 60 °C, then 1N aq. HCl, 75%; (<b>c</b>) <b>13a</b> (X=O, R=2-F), HATU, DIEA, DMF; (<b>d</b>) 4-chlorophenol, DBU, DMF, 70 °C; (<b>e</b>) NaOH, THF, H<sub>2</sub>O; (<b>f</b>) CDI, MeCN, 0 °C to r.t.; (<b>g</b>) <b>13a</b> (X=O, R=2-F), MeCN, 0 °C to r.t.</p>
Full article ">Scheme 3
<p>Synthesis of analogs. Reagents and Conditions: (<b>a</b>) 2-fluorophenol, K<sub>2</sub>CO<sub>3</sub>, CH<sub>3</sub>CN, µwave, 80 °C, 30 min; (<b>b</b>) H<sub>2</sub>, Pt(sulfided)/C, MeOH, 70 °C, H-Cube<sup>®</sup>; (<b>c</b>) 3-(4-chlorophenyl)propanoic acid, HATU, DIEA, DMF.</p>
Full article ">
10 pages, 2063 KiB  
Article
Size Dependence of the Tetragonal to Orthorhombic Phase Transition of Ammonia Borane in Nanoconfinement
by Shah Najiba, Jiuhua Chen, Mohammad S. Islam, Yongzhou Sun, Andriy Durygin and Vadym Drozd
Materials 2024, 17(22), 5672; https://doi.org/10.3390/ma17225672 - 20 Nov 2024
Viewed by 215
Abstract
We have investigated the thermodynamic property modification of ammonia borane via nanoconfinement. Two different mesoporous silica scaffolds, SBA-15 and MCM-41, were used to confine ammonia borane. Using in situ Raman spectroscopy, we examined how pore size influences the phase transition temperature from tetragonal [...] Read more.
We have investigated the thermodynamic property modification of ammonia borane via nanoconfinement. Two different mesoporous silica scaffolds, SBA-15 and MCM-41, were used to confine ammonia borane. Using in situ Raman spectroscopy, we examined how pore size influences the phase transition temperature from tetragonal (I4mm) to orthorhombic (Pmn21) for ammonia borane. In bulk ammonia borane, the phase transition occurs at around 217 K; however, confinement in SBA-15 (with ~8 nm pore sizes) reduces this temperature to approximately 195 K, while confinement in MCM-41 (with pore sizes of 2.1–2.7 nm) further lowers it to below 90 K. This suppression of the phase transition as a function of pore size has not been previously studied using Raman spectroscopy. The stability of the I4mm phase at a much lower temperature can be interpreted by incorporating the surface energy terms to the overall free energy of the system in a simple thermodynamic model, which leads to a significant increase in the surface energy when transitioning from the tetragonal phase to the orthorhombic phase. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Selected Raman spectra of neat NH<sub>3</sub>BH<sub>3</sub> at different temperatures in the spectral region of (<b>a</b>) 600–1650 cm<sup>−1</sup>, (<b>b</b>) 2100–2500 cm<sup>−1</sup> and (<b>c</b>) 3100–3400 cm<sup>−1</sup>. The intense Raman peak of the diamond anvil is truncated in the region of 1270–1500 cm<sup>−1</sup>. The sample temperature is indicated in each spectrum. Arrows indicate peak slitting in the spectra of 217 K and 90 K.</p>
Full article ">Figure 2
<p>Selected Raman spectra of SBA-15/NH<sub>3</sub>BH<sub>3</sub> at different temperatures in the spectral region of (<b>a</b>) 600–1650 cm<sup>−1</sup>, (<b>b</b>) 2100–2500 cm<sup>−1</sup> and (<b>c</b>) 3100–3400 cm<sup>−1</sup>. The intense Raman mode of the diamond anvil is truncated in the region of 1270–1500 cm<sup>−1</sup>. The sample temperature is indicated in each spectrum. Arrows indicate peak slitting in the spectra of 195 K.</p>
Full article ">Figure 3
<p>Selected Raman spectra of MCM-41/NH<sub>3</sub>BH<sub>3</sub> at different temperatures in the spectral region of (<b>a</b>) 600–1650 cm<sup>−1</sup>, (<b>b</b>) 2100–2500 cm<sup>−1</sup> and (<b>c</b>) 3100–3400 cm<sup>−1</sup>. The intense Raman mode of the diamond anvil is truncated in the region of 1270–1500 cm<sup>−1</sup>. The sample temperature is indicated in each spectrum.</p>
Full article ">
23 pages, 7207 KiB  
Article
Water-Soluble Polyglycidol-Grafted Ladder Calix Resorcinarene Oligomers with Open Chain and Cyclic Topologies: Synthesis, Characteristics, and Biological Evaluation
by Hristo Penchev, Erik Dimitrov, Christo Novakov, Emi Haladjova, Ralitsa Veleva, Veselina Moskova-Doumanova, Tanya Topouzova-Hristova and Stanislav Rangelov
Polymers 2024, 16(22), 3219; https://doi.org/10.3390/polym16223219 - 20 Nov 2024
Viewed by 366
Abstract
Ladder oligomers containing calixarene skeletons in the main chain—calix[4]resorcinarene (CRA) ladder macromolecules with open chain and cyclic macromolecules with double ring-like (Noria-type) topologies—bring particular research attention as functional materials with various applications. However, there is still a remarkable lack of studies into the [...] Read more.
Ladder oligomers containing calixarene skeletons in the main chain—calix[4]resorcinarene (CRA) ladder macromolecules with open chain and cyclic macromolecules with double ring-like (Noria-type) topologies—bring particular research attention as functional materials with various applications. However, there is still a remarkable lack of studies into the synthesis of fully water-soluble derivatives of these interesting macromolecules. Research on this topic would allow their bio-based research and application niche to be at least revealed. In the present study, a strategy for the synthesis of water-soluble polyglycidol-derivatized calix resorcinarene ladder oligomers with open chain and cyclic structures is introduced. A grafting from approach was used to build branched or linear polyglycidol chains from the ladder scaffolds. The novel structures were synthesized in quantitative yields and fully characterized by NMR, FTIR and UV–vis spectroscopy, gel permeation chromatography, MALDI-TOF mass spectrometry, analytical ultracentrifugation, and static light scattering to obtain the molar mass characteristics and composition. The biocompatibility and toxicity of the two polyglycidol-derivatized oligomers were investigated and the concentration dependence of the survival of three cell lines of human origin determined. The selective apoptosis effect at relatively low dissolve concentrations toward two kinds of cancerous cell lines was found. Full article
Show Figures

Figure 1

Figure 1
<p>MALDI-TOF spectra of CRA (<b>a</b>) and Noria (<b>b</b>) oligomers.</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR spectra CRA ladder (<b>a</b>) and Noria (<b>b</b>) oligomers in DMSO-d6.</p>
Full article ">Figure 3
<p>GPC trace of Noria taken in THF.</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectrum of CRA-brPG in D<sub>2</sub>O.</p>
Full article ">Figure 5
<p>Sedimentation coefficient distribution from AUC measurements for CRA-brPG. The inset shows a scale-up in the region for s &gt; 4.5 S. The corresponding molar mass values for each peak are shown in <a href="#polymers-16-03219-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 6
<p>Berry plot of CRA-brPG in dilute aqueous solution at 25 °C: experimental point (open symbols); extrapolated points (closed symbols).</p>
Full article ">Figure 7
<p>GPC traces tracked with an UV detector of the precursor Noria-linPEEGE taken in THF.</p>
Full article ">Figure 8
<p><sup>1</sup>H NMR spectra of Noria-PEEGE precursor in CDCl<sub>3</sub> (<b>a</b>) and resulting Noria-linPG in DMSO-d6 (<b>b</b>).</p>
Full article ">Figure 9
<p>MALDI-TOF spectrum of Noria-linPG. The inset shows the difference between the peaks of 74 g/mol, which corresponds to the molar mass of the glycidol unit building the PG arms.</p>
Full article ">Figure 10
<p>Sedimentation coefficient distribution from AUC measurements for Noria-linPG. The corresponding molar mass is shown in <a href="#polymers-16-03219-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 11
<p>Berry plot of Noria-linPG in dilute aqueous solution at 25 °C: experimental point (open symbols); extrapolated points (closed symbols).</p>
Full article ">Figure 12
<p>Viability of human cells 24 (<b>a</b>) and 72 (<b>b</b>) hours after treatment with polyglycidol-derivatized ladder oligomers CRA-brPG and Noria-linPG at a concentration of 5–300 µg/mL. The survival of control cells was taken as 100%.</p>
Full article ">Figure 13
<p>Induction of apoptosis after 72 h treatment with Noria-linPG and CRA-brPG.</p>
Full article ">Scheme 1
<p>Synthesis of ladder CRA (<b>a</b>) and Noria (<b>b</b>) oligomers.</p>
Full article ">Scheme 2
<p>Schematic presentation of the molecular topology of polyglycidol-derivatized CRA (<b>a</b>) and Noria (<b>b</b>) oligomers.</p>
Full article ">Scheme 3
<p>Synthesis of CRA-brPG.</p>
Full article ">Scheme 4
<p>Synthesis of Noria-linPG.</p>
Full article ">
18 pages, 12806 KiB  
Article
3D Printing and Property of Biomimetic Hydroxyapatite Scaffold
by Xueni Zhao, Lingna Li, Yu Zhang, Zhaoyang Liu, Haotian Xing and Zexin Gu
Biomimetics 2024, 9(11), 714; https://doi.org/10.3390/biomimetics9110714 - 20 Nov 2024
Viewed by 310
Abstract
The 3D printing of a biomimetic scaffold with a high hydroxyapatite (HA) content (>80%) and excellent mechanical property is a serious challenge because of the difficulty of forming and printing, insufficient cohesion, and low mechanical property of the scaffold. In this study, hydroxyapatite [...] Read more.
The 3D printing of a biomimetic scaffold with a high hydroxyapatite (HA) content (>80%) and excellent mechanical property is a serious challenge because of the difficulty of forming and printing, insufficient cohesion, and low mechanical property of the scaffold. In this study, hydroxyapatite whiskers (HAWs), with their superior mechanical property, biodegradability, and biocompatibility, were used to reinforce spherical HA scaffolds by 3D printing. The compressive strength and energy absorption capacity of HAW-reinforced spherical HA (HAW/HA) scaffolds increased when the HAW/HA ratio increased from 0:10 to 4:6 and then dropped with any further increases in the HAW/HA ratio. Bioceramic content (HAWs and spherical HA) in the scaffolds reached 83%, and the scaffold with a HAW/HA ratio of 4:6 (4-HAW/HA) exhibited an optimum compressive strength and energy absorption capacity. The scaffold using polyvinyl alcohol (PVA) as an additive possessed a good bonding between HA and PVA as well as a higher strength, which allowed the scaffold to bear a higher stress at the same strain. The compressive strength and toughness of the 4-HAW/HA-PVA scaffold were 1.96 and 1.63 times that of the 4-HAW/HA scaffold with hydroxypropyl methyl cellulose (HPMC), respectively. The mechanical property and inorganic components of the biomimetic HAW/HA scaffold were similar to those of human bone, which would make it ideal for repairing bone defects. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of HAWs.</p>
Full article ">Figure 2
<p>Morphology of (<b>A</b>) 0-HAW/HA, (<b>B</b>) 1-HAW/HA, (<b>C</b>) 2-HAW/HA, (<b>D</b>) 3-HAW/HA, (<b>E</b>) 4-HAW/HA, and (<b>F</b>) 5-HAW/HA.</p>
Full article ">Figure 3
<p>Surface morphology and roughness of (<b>A</b>) 0-HAW/HA, (<b>B</b>) 1-HAW/HA, (<b>C</b>) 2-HAW/HA, (<b>D</b>) 3-HAW/HA, (<b>E</b>) 4-HAW/HA, and (<b>F</b>) 5-HAW/HA.</p>
Full article ">Figure 4
<p>SEM images of (<b>A</b>,<b>B</b>) 0-HAW/HA, (<b>C</b>,<b>D</b>) 1-HAW/HA, (<b>E</b>,<b>F</b>) 2-HAW/HA, (<b>G</b>,<b>H</b>) 3-HAW/HA, (<b>I</b>,<b>J</b>) 4-HAW/HA, and (<b>K</b>,<b>L</b>) 5-HAW/HA.</p>
Full article ">Figure 5
<p>(<b>A</b>) Compressive stress–strain curve of HAW/HA scaffolds and (<b>B</b>) compressive strength of HAW/HA scaffolds with different HAW content. (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 6
<p>Cross-sectional morphologies of (<b>A</b>,<b>B</b>) 0-HAW/HA, (<b>C</b>,<b>D</b>) 1-HAW/HA, (<b>E</b>,<b>F</b>) 2-HAW/HA, (<b>G</b>,<b>H</b>) 3-HAW/HA, (<b>I</b>,<b>J</b>) 4-HAW/HA, and (<b>K</b>,<b>L</b>) 5-HAW/HA.</p>
Full article ">Figure 7
<p>(<b>A</b>) Rheological property of 4-HAW/HA ink, (<b>B</b>) FTIR, (<b>C</b>) EDS, and (<b>D</b>) XRD pattern of 4-HAW/HA scaffold. Vertical lines represent HA standards according to JCPDS # 72-1243.</p>
Full article ">Figure 8
<p>(<b>A</b>,<b>B</b>) SEM images, (<b>C</b>) rheological property, (<b>D</b>) FTIR, (<b>E</b>) EDS, and (<b>F</b>) XRD pattern of 4-HAW/HA-PVA scaffold. The yellow arrows represents the bonding location of PVA, HAWS and spherical HA. Vertical lines represent HA standards according to JCPDS # 72-1243.</p>
Full article ">Figure 9
<p>(<b>A</b>) Compressive stress–strain curves, (<b>B</b>) compressive strength of 4-HAW/HA and 4-HAW/HA-PVA scaffolds (** <span class="html-italic">p</span> &lt; 0.01), (<b>C</b>,<b>D</b>) cross-sectional morphologies of 4 -HAW/HA-PVA, (<b>E</b>) the reinforcing toughing mechanism of PVA and hydroxyapatite bonding, (<b>F</b>) crack deflection and whisker bridging, and (<b>G</b>) whisker pullout. In the figure (<b>E</b>–<b>G</b>), the gray represents the 4 -HAW/HA-PVA matrix, the blue dot represents spherical HA, the yellow bars represents HA whisker.</p>
Full article ">
19 pages, 3701 KiB  
Article
Enhancing Antileishmanial Activity of Amidoxime-Based Compounds Bearing a 4,5-Dihydrofuran Scaffold: In Vitro Screening Against Leishmania amazonensis
by Fabiana Maia Santos Urbancg Moncorvo, Oscar Leonardo Avendaño Leon, Christophe Curti, Youssef Kabri, Sébastien Redon, Eduardo Caio Torres-Santos and Patrice Vanelle
Molecules 2024, 29(22), 5469; https://doi.org/10.3390/molecules29225469 - 20 Nov 2024
Viewed by 246
Abstract
Leishmaniasis, a protozoan disease affecting humans, exposes significant shortcomings in current treatments. In continuation to our previous findings on amidoxime-based antileishmanial compounds bearing a 4,5-dihydrofuran scaffold, twelve new amidoxime derivatives substituted at position 3 with an amide bearing a nitrogen heterocycle were synthesized. [...] Read more.
Leishmaniasis, a protozoan disease affecting humans, exposes significant shortcomings in current treatments. In continuation to our previous findings on amidoxime-based antileishmanial compounds bearing a 4,5-dihydrofuran scaffold, twelve new amidoxime derivatives substituted at position 3 with an amide bearing a nitrogen heterocycle were synthesized. This series was designed to replace the sulfone and aryl group on a previously reported HIT. The synthesis of these compounds involved the following three-step pathway: manganese (III) acetate-based cyclization of a β-ketoester, followed by amidation with LiHMDS and a final reaction with hydroxylamine. Three of them, containing either bromine, chlorine, or methyl substitutions and featuring a pyridine moiety, showed an interesting toxicity–activity relationship in vitro. They exhibited IC50 values of 15.0 µM, 16.0 µM, and 17.0 µM against the promastigote form of the parasite and IC50 values of 0.5 µM, 0.6 µM, and 0.3 µM against the intracellular amastigote form, respectively. A selectivity index (SI) greater than 300 was established between the cytotoxic concentrations (in murine macrophages) and the effective concentrations (against the intracellular form of Leishmania amazonensis). This SI is at least seventy times higher than that observed for Pentamidine and twenty-five times higher than that observed for the reference HIT, as previously reported. Full article
Show Figures

Scheme 1

Scheme 1
<p>Optimization of 4,5-dihydrofuran derivatives bearing an amidoxime group against <span class="html-italic">L. amazonensis</span>.</p>
Full article ">Scheme 2
<p>Synthetic pathways of antileishmanial compounds bearing the 4,5-dihydrofuran-3-carboxamide scaffold and an amidoxime moiety. Reagent and conditions: (i) Mn(OAc)<sub>3</sub> (2.1 equiv.), Cu(OAc)<sub>2</sub> (1 equiv.), 2-methyl-3-phenyl-1-propene (2 equiv.), AcOH, MW, 80 °C, 100 W, and 3 h. (ii) Amine (1.5 equiv.), LiHMDS (2.5 equiv.), Toluene, 70 °C, 15 h, and N<sub>2</sub>. (iii) KO<span class="html-italic">t</span>Bu (10 equiv.), NH<sub>2</sub>OH.HCl (10 equiv.), DMSO, 18 h, 0 °C—RT, and N<sub>2</sub>.</p>
Full article ">Scheme 3
<p>Electron transfer to generate radicals.</p>
Full article ">
15 pages, 1305 KiB  
Review
Morphology and Applications of Self-Assembled Peptide Nucleic Acids
by Luca Domenico D’Andrea and Alessandra Romanelli
Int. J. Mol. Sci. 2024, 25(22), 12435; https://doi.org/10.3390/ijms252212435 - 19 Nov 2024
Viewed by 238
Abstract
Obtaining new materials by exploiting the self-assembly of biomolecules is a very challenging field. In recent years, short peptides and nucleic acids have been used as scaffolds to produce supramolecular structures for different applications in the biomedical and technological fields. In this review, [...] Read more.
Obtaining new materials by exploiting the self-assembly of biomolecules is a very challenging field. In recent years, short peptides and nucleic acids have been used as scaffolds to produce supramolecular structures for different applications in the biomedical and technological fields. In this review, we will focus on the self-assembly of peptide nucleic acids (PNAs), their conjugates with peptides, or other molecules. We will describe the physical properties of the assembled systems and, where described, the application they were designed for. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structure of selected PNAs assembling to produce fibers: (<b>A</b>) R = H, X = 2 [<a href="#B22-ijms-25-12435" class="html-bibr">22</a>]; R= H or CH<sub>2</sub>OH, x= 12 [<a href="#B23-ijms-25-12435" class="html-bibr">23</a>]; (<b>B</b>) y = 1,2 and z = 1,2 [<a href="#B24-ijms-25-12435" class="html-bibr">24</a>,<a href="#B25-ijms-25-12435" class="html-bibr">25</a>]; (<b>C</b>) k = 10 [<a href="#B26-ijms-25-12435" class="html-bibr">26</a>].</p>
Full article ">Figure 2
<p>Chemical structure of selected PNAs assembling to produce spheres: (<b>A</b>) [<a href="#B27-ijms-25-12435" class="html-bibr">27</a>]; (<b>B</b>,<b>D</b>) [<a href="#B28-ijms-25-12435" class="html-bibr">28</a>]; (<b>C</b>) [<a href="#B29-ijms-25-12435" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>Chemical structure of selected PNAs assembling to produce micelles: (<b>A</b>) [<a href="#B30-ijms-25-12435" class="html-bibr">30</a>]; (<b>B</b>) [<a href="#B31-ijms-25-12435" class="html-bibr">31</a>,<a href="#B32-ijms-25-12435" class="html-bibr">32</a>]; (<b>C</b>) [<a href="#B33-ijms-25-12435" class="html-bibr">33</a>].</p>
Full article ">
18 pages, 2519 KiB  
Article
A Tissue-Engineered Construct Based on a Decellularized Scaffold and the Islets of Langerhans: A Streptozotocin-Induced Diabetic Model
by Victor I. Sevastianov, Anna S. Ponomareva, Natalia V. Baranova, Aleksandra D. Belova, Lyudmila A. Kirsanova, Alla O. Nikolskaya, Eugenia G. Kuznetsova, Elizaveta O. Chuykova, Nikolay N. Skaletskiy, Galina N. Skaletskaya, Evgeniy A. Nemets, Yulia B. Basok and Sergey V. Gautier
Life 2024, 14(11), 1505; https://doi.org/10.3390/life14111505 - 19 Nov 2024
Viewed by 353
Abstract
Producing a tissue-engineered pancreas based on a tissue-specific scaffold from a decellularized pancreas, imitating the natural pancreatic tissue microenvironment and the islets of Langerhans, is one of the approaches to treating patients with type 1 diabetes mellitus (T1DM). The aim of this work [...] Read more.
Producing a tissue-engineered pancreas based on a tissue-specific scaffold from a decellularized pancreas, imitating the natural pancreatic tissue microenvironment and the islets of Langerhans, is one of the approaches to treating patients with type 1 diabetes mellitus (T1DM). The aim of this work was to investigate the ability of a fine-dispersed tissue-specific scaffold (DP scaffold) from decellularized human pancreas fragments to support the islets’ survival and insulin-producing function when injected in a streptozotocin-induced diabetic rat model. The developed decellularization protocol allows us to obtain a scaffold with a low DNA content (33 [26; 38] ng/mg of tissue, p < 0.05) and with the preservation of GAGs (0.92 [0.84; 1.16] µg/mg, p < 0.05) and fibrillar collagen (273.7 [241.2; 303.0] µg/mg, p < 0.05). Rat islets of Langerhans were seeded in the obtained scaffolds. The rats with stable T1DM were treated by intraperitoneal injections of rat islets alone and islets seeded on the DP scaffold. The blood glucose level was determined for 10 weeks with a histological examination of experimental animals’ pancreas. A more pronounced decrease in the recipient rats’ glycemia was detected after comparing the islets seeded on the DP scaffold with the control injection (by 71.4% and 51.2%, respectively). It has been shown that the DP scaffold facilitates a longer survival and the efficient function of pancreatic islets in vivo and can be used to engineer a pancreas. Full article
(This article belongs to the Section Cell Biology and Tissue Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Research design.</p>
Full article ">Figure 2
<p>The DP-scaffold characterization. (<b>a</b>–<b>f</b>) A histological picture of a tissue-specific scaffold from a decellularized pancreas (DP scaffold); (<b>a</b>) H&amp;E staining; (<b>b</b>) Masson’s staining; (<b>c</b>) alcian blue staining; (<b>d</b>) immunohistochemical staining with antibodies to type I collagen; (<b>e</b>) staining with orcein for elastic fibers; (<b>f</b>) DAPI staining (scale bar = 100 µm); (<b>g</b>) GAG content; (<b>h</b>) collagen content; (<b>i</b>) DNA content; (<b>j</b>–<b>o</b>) a DP-scaffold cytotoxicity study in vitro; (<b>j</b>,<b>m</b>) negative control; (<b>k</b>,<b>n</b>) positive control; (<b>l</b>,<b>o</b>) a DP scaffold on the surface of the NIH/3T3 monolayer; (<b>j</b>–<b>l</b>) phase contrast microscopy; (<b>m</b>–<b>o</b>) fluorescence LIVE/DEAD; scale bar = 100 µm.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) Isolated healthy rat islets of Langerhans: (<b>a</b>) phase contrast microscopy; (<b>b</b>) dithizone staining; (<b>c</b>) islets cultured for 24 h, acridine orange and propidium iodide (AO/PI) fluorescent staining; (<b>d</b>) functional activity of isolated healthy rat islets of Langerhans cultured for 24 h; (<b>e</b>,<b>f</b>) islets seeded on DP scaffold: (<b>e</b>) phase contrast microscopy; (<b>f</b>) AO/PI fluorescent staining; scale bar = 100 µm.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>c</b>) A pancreas of a healthy rat; (<b>d</b>–<b>f</b>) a pancreas of a rat from the control group with streptozotocin-induced T1DM, 0 week; (<b>g</b>–<b>i</b>) a pancreas of a rat from the control group with streptozotocin-induced T1DM, 10 weeks; (<b>j</b>–<b>l</b>) a pancreas of a rat from experimental group 1 with streptozotocin-induced T1DM after an intraperitoneal injection of the islets of Langerhans; (<b>m</b>–<b>o</b>) a pancreas of a rat from experimental group 2 with streptozotocin-induced T1DM after an intraperitoneal injection of the islets of Langerhans with a scaffold; (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>) H&amp;E staining; (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>) insulin immunohistochemical staining; (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>) glucagon immunohistochemical staining; scale bar = 100 µm.</p>
Full article ">Figure 5
<p>The dynamics of blood glucose levels in rats with streptozotocin-induced T1DM from the control (without treatment) and experimental groups after an intraperitoneal injection of the islets of Langerhans or the islets of Langerhans seeded on a DP scaffold. <sup>∆</sup> <span class="html-italic">p</span> &lt; 0.05: control group vs. experimental group 1; <sup>#</sup> <span class="html-italic">p</span> &lt; 0.05: control group vs. experimental group 2; * <span class="html-italic">p</span> &lt; 0.05: experimental group 1 vs. experimental group 2.</p>
Full article ">
29 pages, 1081 KiB  
Review
Hydrogel-Enhanced Autologous Chondrocyte Implantation for Cartilage Regeneration—An Update on Preclinical Studies
by Xenab Ahmadpoor, Jessie Sun, Nerone Douglas, Weimin Zhu and Hang Lin
Bioengineering 2024, 11(11), 1164; https://doi.org/10.3390/bioengineering11111164 - 19 Nov 2024
Viewed by 335
Abstract
Autologous chondrocyte implantation (ACI) and matrix-induced ACI (MACI) have demonstrated improved clinical outcomes and reduced revision rates for treating osteochondral and chondral defects. However, their ability to achieve lasting, fully functional repair remains limited. To overcome these challenges, scaffold-enhanced ACI, particularly utilizing hydrogel-based [...] Read more.
Autologous chondrocyte implantation (ACI) and matrix-induced ACI (MACI) have demonstrated improved clinical outcomes and reduced revision rates for treating osteochondral and chondral defects. However, their ability to achieve lasting, fully functional repair remains limited. To overcome these challenges, scaffold-enhanced ACI, particularly utilizing hydrogel-based biomaterials, has emerged as an innovative strategy. These biomaterials are intended to mimic the biological composition, structural organization, and biomechanical properties of native articular cartilage. This review aims to provide comprehensive and up-to-date information on advancements in hydrogel-enhanced ACI from the past decade. We begin with a brief introduction to cartilage biology, mechanisms of cartilage injury, and the evolution of surgical techniques, particularly looking at ACI. Subsequently, we review the diversity of hydrogel scaffolds currently undergoing development and evaluation in preclinical studies for articular cartilage regeneration, emphasizing chondrocyte-laden hydrogels applicable to ACI. Finally, we address the key challenges impeding effective clinical translation, with particular attention to issues surrounding fixation and integration, aiming to inform and guide the future progression of tissue engineering strategies. Full article
(This article belongs to the Section Regenerative Engineering)
Show Figures

Figure 1

Figure 1
<p>Different types of cartilage injury.</p>
Full article ">Figure 2
<p>A successful ACI-based cartilage repair needs to have efficacious fixation and integration of hydrogels and implants.</p>
Full article ">
Back to TopTop