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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Search Results (403)

Search Parameters:
Keywords = sol gel transition

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 8753 KiB  
Article
Effect of TiO2 Coating on Structure and Electrochemical Performance of LiNi0.6Co0.2Mn0.2O2 Cathode Material for Lithium-Ion Batteries
by Lin Li, Zhongyu Li, Zhifan Kuang, Hao Zheng, Minjian Yang, Jianwen Liu, Shiquan Wang and Hongying Liu
Materials 2024, 17(24), 6222; https://doi.org/10.3390/ma17246222 (registering DOI) - 19 Dec 2024
Abstract
High-nickel ternary LiNi0.6Co0.2Mn0.2O2 (NCM622) is a promising cathode material for lithium-ion batteries due to its high discharge-specific capacity and energy density. However, problems of NCM622 materials, such as unstable surface structure, lithium–nickel co-segregation, and intergranular cracking, [...] Read more.
High-nickel ternary LiNi0.6Co0.2Mn0.2O2 (NCM622) is a promising cathode material for lithium-ion batteries due to its high discharge-specific capacity and energy density. However, problems of NCM622 materials, such as unstable surface structure, lithium–nickel co-segregation, and intergranular cracking, led to a decrease in the cycling performance of the material and an inability to fully utilize high specific capacity. Surface coating was the primary approach to address these problems. The effect of TiO2 coating prepared by the sol–gel method on the performance of LiNi0.6Co0.2Mn0.2O2 was studied, mainly including the morphology, cell structure, and electrochemical properties. LiNi0.6Co0.2Mn0.2O2 was coated by TiO2 with a thickness of about 5 nm. Compared with the pristine NCM622 electrode, the electrochemical performance of the TiO2-coated NCM622 electrodes is improved. Among all TiO2-coated NCM622, the NCM622 cathode with TiO2 coating content of 0.5% demonstrates the highest capacity retention of 89.3% and a discharge capacity of 163.9 mAh g−1, in contrast to 80.9% and145 mAh g−1 for the pristine NCM622 electrode, after 100 cycles at 0.3 C between 3 and 4.3 V. The cycle life of the 5 wt% TiO2-coated NCM622 electrode is significantly improved at a high cutoff voltage of 4.6 V. The significantly enhanced cycling performance of TiO2-coated NCM622 materials could be attributed to the TiO2 coating layer that could block the contact between the material surface and the electrolyte, reducing the interface side reaction and inhibiting the transition metal dissolution. At the same time, the coating layer maintained the stability of layered structures, thus reducing the polarization phenomenon of the electrode and alleviating the irreversible capacity loss in the cycle process. Full article
(This article belongs to the Special Issue Catalytic Materials and Renewable Chemistry for Energy and Fuels)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of synthesis procedure for TiO<sub>2</sub>-NCM622.</p>
Full article ">Figure 2
<p>XRD patterns of NCM622 and TiO<sub>2</sub> coated NCM samples.</p>
Full article ">Figure 3
<p>The XPS spectra of 5 wt% TiO<sub>2</sub>-NCM: (<b>a</b>) survey spectrum, (<b>b</b>) C 1s, (<b>c</b>) O 1s, (<b>d</b>) Ni 2p, (<b>e</b>) Co 2p, (<b>f</b>) Mn 2p, (<b>g</b>) Ti 2p.</p>
Full article ">Figure 3 Cont.
<p>The XPS spectra of 5 wt% TiO<sub>2</sub>-NCM: (<b>a</b>) survey spectrum, (<b>b</b>) C 1s, (<b>c</b>) O 1s, (<b>d</b>) Ni 2p, (<b>e</b>) Co 2p, (<b>f</b>) Mn 2p, (<b>g</b>) Ti 2p.</p>
Full article ">Figure 4
<p>The SEM images of (<b>a</b>) NCM622, (<b>b</b>) 2 wt% TiO<sub>2</sub>-NCM, (<b>c</b>) 5 wt% TiO<sub>2</sub>-NCM, and (<b>d</b>) 8 wt% TiO<sub>2</sub>-NCM. (<b>e</b>) TEM image of 5 wt% TiO<sub>2</sub>-NCM; (<b>f</b>) HRTEM image; (<b>g</b>) FE-SEM elemental mapping of 5 wt% TiO<sub>2</sub>-NCM microspheres. (<b>h</b>,<b>i</b>) HRTEM images of 5 wt% TiO<sub>2</sub>-NCM.</p>
Full article ">Figure 5
<p>CV curves of 5 wt% TiO<sub>2</sub>-NCM.</p>
Full article ">Figure 6
<p>The charge-discharge curves of (<b>a</b>) NCM622, (<b>b</b>) 2 wt% TiO<sub>2</sub>-NCM, (<b>c</b>) 5 wt% TiO<sub>2</sub>-NCM, and (<b>d</b>) 8 wt% TiO<sub>2</sub>-NCM samples. (<b>e</b>) The long-term cycle of NCM, 2 wt% TiO<sub>2</sub>-NCM, 5 wt% TiO<sub>2</sub>-NCM, and 8 wt% TiO<sub>2</sub>-NCM samples at 0.3 C. (<b>f</b>) Rate performance of NCM, 2 wt% TiO<sub>2</sub>-NCM, 5 wt% TiO<sub>2</sub>-NCM, and 8 wt% TiO<sub>2</sub>-NCM samples at different current densities.</p>
Full article ">Figure 7
<p>The charge-discharge curves of (<b>a</b>) NCM622, (<b>b</b>) 5 wt% TiO<sub>2</sub>-NCM sample, and (<b>c</b>) cyclic curve under 0.5 C and 4.6 V conditions.</p>
Full article ">Figure 8
<p>(<b>a</b>) Nyquist plots of NCM, 2 wt% TiO<sub>2</sub>-NCM, 5 wt% TiO<sub>2</sub>-NCM, and 8 wt% TiO<sub>2</sub>-NCM electrodes (the inset is an equivalent circuit.); (<b>b</b>) the linear fitting diagram of the Warburg impedance.</p>
Full article ">
21 pages, 8687 KiB  
Article
Development and Characterization of Dual-Loaded Niosomal Ion-Sensitive In Situ Gel for Ocular Delivery
by Viliana Gugleva, Rositsa Mihaylova, Katya Kamenova, Dimitrina Zheleva-Dimitrova, Denitsa Stefanova, Virginia Tzankova, Maya Margaritova Zaharieva, Hristo Najdenski, Aleksander Forys, Barbara Trzebicka, Petar D. Petrov and Denitsa Momekova
Gels 2024, 10(12), 816; https://doi.org/10.3390/gels10120816 - 11 Dec 2024
Viewed by 477
Abstract
The study investigates the development and characterization of dual-loaded niosomes incorporated into ion-sensitive in situ gel as a potential drug delivery platform for ophthalmic application. Cannabidiol (CBD) and epigallocatechin-3-gallate (EGCG) simultaneously loaded niosomes were prepared via the thin film hydration (TFH) method followed [...] Read more.
The study investigates the development and characterization of dual-loaded niosomes incorporated into ion-sensitive in situ gel as a potential drug delivery platform for ophthalmic application. Cannabidiol (CBD) and epigallocatechin-3-gallate (EGCG) simultaneously loaded niosomes were prepared via the thin film hydration (TFH) method followed by pulsatile sonication and were subjected to comprehensive physicochemical evaluation. The optimal composition was included in a gellan gum-based in situ gel, and the antimicrobial activity, in vitro toxicity in a suitable corneal epithelial model (HaCaT cell line), and antioxidant potential of the hybrid system were further assessed. Dual-loaded niosomes based on Span 60, Tween 60, and cholesterol (3.5:3.5:3 mol/mol) were characterized by appropriate size (250 nm), high entrapment efficiency values for both compounds (85% for CBD and 50% for EGCG) and sustained release profiles. The developed hybrid in situ gel exhibited suitable rheological characteristics to enhance the residence time on the ocular surface. The conducted microbiological studies reveal superior inhibition of methicillin-resistant Staphylococcus aureus (MRSA) adhesion by means of the niosomal in situ gel compared to the blank gel and untreated control. Regarding the antioxidant potential, the dual loading of CBD and EGCG in niosomes enhances their protective properties, and the inclusion of niosomes in gel form preserves these effects. The obtained outcomes indicate the developed niosomal in situ gel as a promising drug delivery platform in ophthalmology. Full article
(This article belongs to the Special Issue Composite Hydrogels for Biomedical Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Size distributions of empty and drug-loaded niosomes.</p>
Full article ">Figure 2
<p>Cryo-TEM images of (<b>a</b>) empty niosomes (N6); (<b>b</b>) CBD-loaded niosomes (N5).</p>
Full article ">Figure 3
<p>Viscosity as a function of the shear rate of G6, G6N, and G6N:CBD:EGCG formulations at 25 °C.</p>
Full article ">Figure 4
<p>Variation in elastic (G′) and loss (G″) moduli as a function of shear stress (τ) of G6, G6N, and G6N:CBD:EGCG formulations. All measurements were carried out at 35 °C.</p>
Full article ">Figure 5
<p>In vitro release profile of CBD and EGCG from optimal niosomal formulation (N3) and its hybrid in situ gelling system (G6N:CBD:EGCG). N3:CBD, G6N:CBD denote cannabidiol release from niosomal suspension and hybrid niosomal gel, respectively, whereas N3:EGCG and G6N:EGCG represent the release profiles of epigallocatechin-3-gallate from niosomes and its hybrid niosomal gel formulation. Each value is presented as mean ± SD (n = 3).</p>
Full article ">Figure 6
<p>Quantitative evaluation of MRSA biofilm formation after exposure to blank (G6N) or hybrid niosomal drug-loaded gel G6N:CBD:EGCG (1/0.5 mg/mL). Legend: Co—untreated control; Dilution 1:8 = 0.125/0.06125 mg/mL; Dilution 1:16 = 0.0625/0.03125 mg/mL; Dilution 1:32 = 0.03125/0.0156 mg/mL. Each value is presented as mean ± SD (n = 4).</p>
Full article ">Figure 7
<p>Cytotoxicity on HaCaT cells of: (<b>A</b>) empty niosomes (N6); (<b>B</b>) free cannabidiol (CBD); (<b>C</b>) free epigallocatechin (EGCG); (<b>D</b>) combination of free cannabidiol and free epigallocatechin (CBD + EGCG); (<b>E</b>) dual-loaded CBD and EGCG vesicles (N:CBD:GCG, formulation) niosomes and (<b>F</b>) niosomal in situ gel based on double-loaded niosomes (G6N:CBD:EGCG), measured by MTT assay. All groups were compared statistically vs. untreated controls by one-way ANOVA with Dunnet’s post hoc test. The results are expressed as means ± SD of triplicate assays (n = 8). *** <span class="html-italic">p</span> &lt; 0.001 vs. control (CTRL, untreated control cells).</p>
Full article ">Figure 8
<p>Protective effects of (<b>A</b>) empty niosomes; (<b>B</b>) free CBD; (<b>C</b>) free EGCG; (<b>D</b>) combination of free CBD and free EGCG (CBD + EGCG); (<b>E</b>) dual-loaded CBD and EGCG (N:CBD:EGCG) niosomes and (<b>F</b>) niosomal in situ gel based on double-loaded niosomes (G6N:CBD:EGCG) in a H<sub>2</sub>O<sub>2</sub>-induced damage model in human keratinocyte HaCaT cell line. The results are expressed as means ± SD of triplicate assays (n = 8). ANOVA with Dunnett’s post-test. * <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 vs. H<sub>2</sub>O<sub>2</sub> CTRL (untreated control cells); H<sub>2</sub>O<sub>2</sub> cells treated with H<sub>2</sub>O<sub>2</sub> (200 µM).</p>
Full article ">
14 pages, 3625 KiB  
Article
MnOx and Pd Surface Functionalization of TiO2 Thin Films via Photodeposition UV Dose Control
by Bozhidar I. Stefanov and Hristo G. Kolev
Photochem 2024, 4(4), 474-487; https://doi.org/10.3390/photochem4040029 - 22 Nov 2024
Viewed by 915
Abstract
This study investigated the influence of the ultraviolet (UV) dose (DUV) on the photodeposition of MnOx and Pd cocatalysts on 300-nm-thick anatase TiO2 thin films, which were prepared via sol–gel dip-coating on a glass substrate. MnOx [...] Read more.
This study investigated the influence of the ultraviolet (UV) dose (DUV) on the photodeposition of MnOx and Pd cocatalysts on 300-nm-thick anatase TiO2 thin films, which were prepared via sol–gel dip-coating on a glass substrate. MnOx and Pd were photodeposited using increasing UV doses ranging from 5 to 20 J cm−2, from 5 mM aqueous electrolytes based on Mn2+/IO3 or Pd2+, respectively. The effect of the DUV on the MnOx photodeposition resulted in an increase in Mn2+ surface content, from 2.7 to 5.2 at.%, as determined using X-ray photoelectron spectroscopy (XPS). For Pd, increasing the UV dose led to a reduction in the oxidation state, transitioning from Pd2+ to Pd0, while the overall Pd surface content range remained relatively steady at 2.2–2.4 at.%. Both MnOx/TiO2 and Pd/TiO2 exhibited proportional enhancements in photocatalytic activity towards the degradation of methylene blue. Notably, Pd/TiO2 demonstrated a significant improvement in photocatalytic performance, surpassing that of pristine TiO2. In contrast, TiO2 samples functionalized through wet impregnation and thermal treatment in the same electrolytes showed overall lower photocatalytic activity compared to those functionalized via photodeposition. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the experimental protocol used for TiO<sub>2</sub> thin film functionalization in this study, illustrating the two employed routes: photodeposition and thermal wet chemical (incipient wetness) functionalization.</p>
Full article ">Figure 2
<p>(<b>a</b>) Geometry of the UV photodeposition exposure system. (<b>b</b>) Emission spectrum of the UV LED, with the emission maximum (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="sans-serif">λ</mi> </mrow> <mrow> <mi mathvariant="normal">m</mi> <mi mathvariant="normal">a</mi> <mi mathvariant="normal">x</mi> </mrow> </msup> </mrow> </semantics></math>) indicated. (<b>c</b>) Variation in the UV emission intensity (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">I</mi> </mrow> <mrow> <mi mathvariant="normal">U</mi> <mi mathvariant="normal">V</mi> </mrow> </msup> </mrow> </semantics></math>) and cumulative photodeposition UV dose (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">D</mi> </mrow> <mrow> <mi mathvariant="normal">U</mi> <mi mathvariant="normal">V</mi> </mrow> </msup> </mrow> </semantics></math>) rates over a 45 min exposure period.</p>
Full article ">Figure 3
<p>Schematic representation of (<b>a</b>) the photocatalytic reactor setup used for the methylene blue (MB) photocatalytic oxidation (PCO) experiments and (<b>b</b>) an example of a <math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">C</mi> </mrow> <mrow> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">B</mi> </mrow> </msup> </mrow> </semantics></math> concentration profile obtained during MB PCO experiments with pristine TiO<sub>2</sub> films, along with MB photolysis data from blank experiments. Error bars on the <math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">C</mi> </mrow> <mrow> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">B</mi> </mrow> </msup> </mrow> </semantics></math> profile indicate standard errors based on triplicate measurements, demonstrating the reproducibility of the experimental data.</p>
Full article ">Figure 4
<p>SEM micrographs of (<b>a</b>) the surface morphology and (<b>b</b>) cross-sectional view of the pristine TiO<sub>2</sub> thin film, the surface morphology of the thermally functionalized (<b>c</b>) MnO<sub>x</sub>/TiO<sub>2</sub> and (<b>e</b>) Pd/TiO<sub>2</sub> thin films, and the photodeposition-functionalized (<b>d</b>) MnO<sub>x</sub>/TiO<sub>2</sub> and (<b>f</b>) Pd/TiO<sub>2</sub> catalysts at a UV dose of 20 J cm<sup>−2</sup>.</p>
Full article ">Figure 5
<p>EDX spectra of (<b>a</b>) MnO<sub>x</sub>/TiO<sub>2</sub> and (<b>c</b>) Pd/TiO<sub>2</sub> films, photodeposition-functionalized at 20 J cm<sup>−2</sup>, along with the corresponding element maps for (<b>b</b>) Mn, Ti, and O in the MnO<sub>x</sub>/TiO<sub>2</sub> case and (<b>d</b>) Pd, Ti, and O in the Pd/TiO<sub>2</sub> case.</p>
Full article ">Figure 6
<p>XRD patterns of the pristine TiO<sub>2</sub> and the MnO<sub>x</sub>- and Pd-functionalized TiO<sub>2</sub> layers, with photodeposition functionalization occurring at a UV dose of 20 J cm<sup>−2</sup>.</p>
Full article ">Figure 7
<p>Comparison of Raman spectra for pristine TiO<sub>2</sub> and layers functionalized through thermal treatment and photodeposition (PD) for (<b>a</b>) MnO<sub>x</sub>/TiO<sub>2</sub> and (<b>b</b>) Pd/TiO<sub>2</sub>.</p>
Full article ">Figure 8
<p>XPS spectra for thermally functionalized and photodeposition-functionalized TiO<sub>2</sub> samples, showing (<b>a</b>) the Mn2p region for MnO<sub>x</sub>/TiO<sub>2</sub> and (<b>b</b>) the Pd3d region for Pd/TiO<sub>2</sub>.</p>
Full article ">Figure 9
<p>Influence of MnO<sub>x</sub>/TiO<sub>2</sub> and Pd/TiO<sub>2</sub> functionalization conditions on (<b>a</b>) the saturation coverage of MB (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">θ</mi> </mrow> <mrow> <mi>MB</mi> </mrow> </msup> </mrow> </semantics></math>) and (<b>b</b>) MB PCO rate (<math display="inline"><semantics> <mrow> <msup> <mrow> <mi mathvariant="normal">r</mi> </mrow> <mrow> <mi>MB</mi> </mrow> </msup> </mrow> </semantics></math>). Data obtained from triplicate runs. Error bars signify standard error.</p>
Full article ">
19 pages, 4523 KiB  
Article
Cr3+ Doping Effects on Structural, Optical, and Morphological Characteristics of BaTiO3 Nanoparticles and Their Bioactive Behavior
by Efracio Mamani Flores, Bertha Silvana Vera Barrios, Julio César Huillca Huillca, Jesús Alfredo Chacaltana García, Carlos Armando Polo Bravo, Henry Edgardo Nina Mendoza, Alberto Bacilio Quispe Cohaila, Francisco Gamarra Gómez, Rocío María Tamayo Calderón, Gabriela de Lourdes Fora Quispe and Elisban Juani Sacari Sacari
Crystals 2024, 14(11), 998; https://doi.org/10.3390/cryst14110998 - 19 Nov 2024
Viewed by 682
Abstract
This study investigates the effects of chromium (Cr3+) doping on BaTiO3 nanoparticles synthesized via the sol–gel route. X-ray diffraction confirms a Cr-induced cubic-to-tetragonal phase transition, with lattice parameters and crystallite size varying systematically with Cr3+ content. UV–visible spectroscopy reveals [...] Read more.
This study investigates the effects of chromium (Cr3+) doping on BaTiO3 nanoparticles synthesized via the sol–gel route. X-ray diffraction confirms a Cr-induced cubic-to-tetragonal phase transition, with lattice parameters and crystallite size varying systematically with Cr3+ content. UV–visible spectroscopy reveals a monotonic decrease in bandgap energy from 3.168 eV (pure BaTiO3) to 2.604 eV (5% Cr3+-doped BaTiO3). Raman and FTIR spectroscopy elucidate structural distortions and vibrational mode alterations caused by Cr3+ incorporation. Transmission electron microscopy and energy-dispersive X-ray spectroscopy verify nanoscale morphology and successful Cr3+ doping (up to 1.64 atom%). Antioxidant activity, evaluated using the DPPH assay, shows stable radical scavenging for pure BaTiO3 (40.70–43.33%), with decreased activity at higher Cr3+ doping levels. Antibacterial efficacy against Escherichia coli peaks at 0.5% Cr3+ doping (10.569 mm inhibition zone at 1.5 mg/mL), decreasing at higher concentrations. This study demonstrates the tunability of structural, optical, and bioactive properties in Cr3+-doped BaTiO3 nanoparticles, highlighting their potential as multifunctional materials for electronics, photocatalysis, and biomedical applications. Full article
(This article belongs to the Special Issue Synthesis and Characterization of Oxide Nanoparticles)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Thermogravimetric analysis and (<b>b</b>) differential scanning calorimetry of pure and Cr<sup>3+</sup>-doped BaTiO<sub>3</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>) X-ray diffraction patterns of pure and doped BaTiO<sub>3</sub> and (<b>b</b>) amplification of XRD peaks within the 44–46.5° range.</p>
Full article ">Figure 3
<p>Raman spectra of (<b>a</b>) Pristine BaTiO<sub>3</sub> and (<b>b</b>) Cr<sup>3+</sup>-doped BaTiO<sub>3</sub>.</p>
Full article ">Figure 4
<p>FTIR spectra of pure and Cr<sup>3+</sup>-doped BaTiO<sub>3</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV–visible diffuse reflectance spectrum. (<b>b</b>) Kubelka–Munk plot for bandgap calculation.</p>
Full article ">Figure 6
<p>Photoluminescence spectrums of pure and Cr<sup>3+</sup>-doped BaTiO<sub>3</sub><b>.</b></p>
Full article ">Figure 7
<p>Transmission electron microscopy microphotography of (<b>a</b>) BaTiO<sub>3</sub>, (<b>b</b>) BaTiO<sub>3</sub>-0.3%Cr<sup>3+</sup>, (<b>c</b>) BaTiO<sub>3</sub>-0.5%Cr<sup>3+</sup>, (<b>d</b>) BaTiO<sub>3</sub>-1%Cr<sup>3+</sup>, (<b>e</b>) BaTiO<sub>3</sub>-3%Cr<sup>3+</sup>, and (<b>f</b>) BaTiO<sub>3</sub>-5%Cr<sup>3+</sup>.</p>
Full article ">Figure 8
<p>Antioxidant activity of pure and Cr<sup>3+</sup>-doped BaTiO<sub>3.</sub></p>
Full article ">
20 pages, 5385 KiB  
Article
Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels
by Alan D. Alcantar Mendoza, Antonieta García Murillo, Felipe de J. Carrillo Romo and José Guzmán Mendoza
Gels 2024, 10(11), 736; https://doi.org/10.3390/gels10110736 - 13 Nov 2024
Viewed by 597
Abstract
This study compared the chemical, structural, and luminescent properties of xerogel-based ceramic powders (CPs) with those of a new series of crystallized aerogels (CAs) synthesized by the epoxy-assisted sol–gel process. Materials with different proportions of Eu3+ (2, 5, 8, and 10 mol%) [...] Read more.
This study compared the chemical, structural, and luminescent properties of xerogel-based ceramic powders (CPs) with those of a new series of crystallized aerogels (CAs) synthesized by the epoxy-assisted sol–gel process. Materials with different proportions of Eu3+ (2, 5, 8, and 10 mol%) were synthesized in Lu2O3 host matrices, as well as a Eu2O3 matrix for comparative purposes. The products were analyzed by infrared spectroscopy (IR), X-ray diffraction (XRD), scanning electron microscopy (SEM) with energy-dispersive spectroscopy (EDS), transmission electron microscopy (TEM), photoluminescence analysis, and by the Brunauer–Emmett–Teller (BET) technique. The results show a band associated with the M-O bond, located at around 575 cm−1. XRD enabled us to check two ensembles: matrices (Lu2O3 or Eu2O3) and doping (Lu2O3:Eu3+) with appropriate chemical compositions featuring C-type crystal structures and intense reflections by the (222) plane, with an interplanar distance of around 0.3 nm. Also, the porous morphology presented by the materials consisted of interconnected particles that formed three-dimensional networks. Finally, emission bands due to the energy transitions (5DJ, where J = 0, 1, 2, and 3) were caused by the Eu3+ ions. The samples doped at 10 mol% showed orange-pink photoluminescence and had the longest disintegration times and greatest quantum yields with respect to the crystallized Eu2O3 aerogel. Full article
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>IR spectra of Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup>: (<b>a</b>) ceramic powders and (<b>b</b>) crystallized aerogels.</p>
Full article ">Figure 2
<p>XRD patterns of the (<b>a</b>) ceramic powders, (<b>b</b>) crystallized aerogels of the Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup> system and (<b>c</b>) schematic representation of the crystalline structure of C-type RE<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 3
<p>Crystallite size vs. lattice distortion degree for (<b>a</b>) ceramic powders and (<b>b</b>) crystallized aerogels.</p>
Full article ">Figure 4
<p>SEM images at different magnifications of (<b>a</b>) ×50,000 and (<b>b</b>) ×100,000, and (<b>c</b>) EDS analysis of the CPLu sample. SEM images at different magnifications of (<b>d</b>) ×50,000 and (<b>e</b>) ×100,000, and (<b>f</b>) EDS analysis of the CPLu10 sample.</p>
Full article ">Figure 5
<p>SEM images at different magnifications of (<b>a</b>) ×50,000 and (<b>b</b>) ×100,000, and (<b>c</b>) EDS analysis of the CALu sample. SEM images at different magnifications of (<b>d</b>) ×50,000 and (<b>e</b>) ×100,000, and (<b>f</b>) EDS analysis of the CALu10 sample.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) TEM images of the CPLu10 sample at different magnifications, (<b>c</b>) HRTEM, and (<b>d</b>) SAED analysis; (<b>e</b>,<b>f</b>) TEM images of the CALu10 sample at different magnifications, (<b>g</b>) HRTEM, and (<b>h</b>) SAED analysis.</p>
Full article ">Figure 7
<p>Photoluminescence study results: (<b>a</b>) excitation and (<b>b</b>) emissions of the ceramic powders, and (<b>c</b>) excitation and (<b>d</b>) emissions of the aerogels.</p>
Full article ">Figure 8
<p>Energy-level diagram of the transfer energy of the Eu<sup>3+</sup> in the CPs and CAs with the Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup> system.</p>
Full article ">Figure 9
<p>(<b>a</b>) CIE diagram and (<b>b</b>) results of the decay time and quantum yield of the Eu<sub>2</sub>O<sub>3</sub> matrices and samples, showing strong photoluminescence in the CP and CA samples doped at 10 mol%.</p>
Full article ">Figure 10
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isotherms and (<b>b</b>) BJH pore-size distribution curves for the samples CPLu10 and CALu10.</p>
Full article ">Figure 11
<p>Stages in the sol–gel process used for the Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup> system from metal salt to ceramic powders and crystallized aerogel: (<b>a</b>) metal salt, (<b>b</b>) sol, (<b>c</b>) wet gel, (<b>d</b>) aged monolithic gel, (<b>e</b>) general xerogel of the Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup> system, (<b>f</b>) ceramic powders (the CPLu10 sample) under UV radiation, (<b>g</b>) supercritical drying chamber, (<b>h</b>) general aerogel of the Lu<sub>2</sub>O<sub>3</sub>:Eu<sup>3+</sup> system, and (<b>i</b>) a crystallized aerogel (the CALu10 sample) under UV radiation.</p>
Full article ">
21 pages, 2940 KiB  
Article
Cord Blood Platelet Lysate-Loaded Thermo-Sensitive Hydrogels for Potential Treatment of Chronic Skin Wounds
by Arianna Grivet-Brancot, Marianna Buscemi, Gianluca Ciardelli, Simona Bronco, Susanna Sartori, Claudio Cassino, Tamer Al Kayal, Paola Losi, Giorgio Soldani and Monica Boffito
Pharmaceutics 2024, 16(11), 1438; https://doi.org/10.3390/pharmaceutics16111438 - 11 Nov 2024
Viewed by 654
Abstract
Background/Objectives: Chronic skin wounds (CSWs) are a worldwide healthcare problem with relevant impacts on both patients and healthcare systems. In this context, innovative treatments are needed to improve tissue repair and patient recovery and quality of life. Cord blood platelet lysate (CB-PL) holds [...] Read more.
Background/Objectives: Chronic skin wounds (CSWs) are a worldwide healthcare problem with relevant impacts on both patients and healthcare systems. In this context, innovative treatments are needed to improve tissue repair and patient recovery and quality of life. Cord blood platelet lysate (CB-PL) holds great promise in CSW treatment thanks to its high growth factors and signal molecule content. In this work, thermo-sensitive hydrogels based on an amphiphilic poly(ether urethane) (PEU) were developed as CB-PL carriers for CSW treatment. Methods: A Poloxamer 407®-based PEU was solubilized in aqueous medium (10 and 15% w/v) and added with CB-PL at a final concentration of 20% v/v. Hydrogels were characterized for their gelation potential, rheological properties, and swelling/dissolution behavior in a watery environment. CB-PL release was also tested, and the bioactivity of released CB-PL was evaluated through cell viability, proliferation, and migration assays. Results: PEU aqueous solutions with concentrations in the range 10–15% w/v exhibited quick (within a few minutes) sol-to-gel transition at around 30–37 °C and rheological properties modulated by the PEU concentration. Moreover, CB-PL loading within the gels did not affect the overall gel properties. Stability in aqueous media was dependent on the PEU concentration, and payload release was completed between 7 and 14 days depending on the polymer content. The CB-PL-loaded hydrogels also showed biocompatibility and released CB-PL induced keratinocyte migration and proliferation, with scratch wound recovery similar to the positive control (i.e., CB-PL alone). Conclusions: The developed hydrogels represent promising tools for CSW treatment, with tunable gelation properties and residence time and the ability to encapsulate and deliver active biomolecules with sustained and controlled kinetics. Full article
Show Figures

Figure 1

Figure 1
<p>Rheological characterization of SHP407-based hydrogels with 10% and 15% <span class="html-italic">w</span>/<span class="html-italic">v</span> concentrations (in light blue and red, respectively): (<b>A</b>) trend of viscosity as function of temperature measured during temperature ramp test; (<b>B</b>) trends of storage (G′) and loss (G″) moduli (continuous and dashed lines, respectively) as function of applied deformation measured during strain sweep test at 37 °C; (<b>C</b>,<b>D</b>) G′ and G″ trends (continuous and dashed lines, respectively) as function of angular frequency measured during frequency sweep test at 25 and 37 °C.</p>
Full article ">Figure 2
<p>Rheological characterization of PL-loaded SHP407-based hydrogels. (<b>A</b>,<b>B</b>) Strain sweep tests at 37 °C. (<b>C</b>,<b>D</b>) Temperature ramp tests. (<b>E</b>,<b>F</b>) Frequency sweep tests at 25 °C. (<b>G</b>,<b>H</b>) Frequency sweep tests at 37 °C.</p>
Full article ">Figure 3
<p>Stability tests for PL-loaded SHP407-based hydrogels. Percentages of (<b>A</b>) swelling and (<b>B</b>) dry weight loss measured for SHP407 10% <span class="html-italic">w</span>/<span class="html-italic">v</span>_PL 20% <span class="html-italic">v</span>/<span class="html-italic">v</span> (light blue) and SHP407 15% <span class="html-italic">w</span>/<span class="html-italic">v</span>_PL 20% <span class="html-italic">v</span>/<span class="html-italic">v</span> (red) samples over time. **** = <span class="html-italic">p</span> ≤ 0.0001.</p>
Full article ">Figure 4
<p>(<b>A</b>) CB-PL release profile (as percentages of the initial encapsulated amount) from the developed SHP407-based hydrogels measured through a BCA assay. (<b>B</b>) The PDGF-AB release profile obtained with the ELISA test.</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) Effects of released CB-PL from SHP407 15% <span class="html-italic">w</span>/<span class="html-italic">v</span>_PL 20% <span class="html-italic">v</span>/<span class="html-italic">v</span> gel on viability and proliferation of L929 cells. Viability (<b>A</b>) and proliferation (<b>B</b>) were assessed using MTT assay and BrdU incorporation assay, respectively, following 72 and 48 hours of incubation with CB-PL containing hydrogel extracts collected at different time points (1, 2, 3, and 7 days). Percentages of cell viability and proliferation were calculated vs. complete medium (assumed as 100%). Data are reported as mean ± SD of values obtained from three independent experiments with three replicates each; * <span class="html-italic">p</span> &lt; 0.05 vs. serum-free medium. (<b>C</b>) Phase-contrast micrographs of scratch test. Scratch wound closure was evaluated after 20 h of exposure to CB-PL containing hydrogel extracts collected at different time points (1, 2, 3, and 7 days), CB-PL 2% serum-free medium and serum-free medium (positive and negative controls, respectively). Scale bar = 500 μm.</p>
Full article ">
11 pages, 5031 KiB  
Article
Transition Metal-Doped Layered Iron Vanadate (FeV3-xMxO9.2.6H2O, M = Co, Mn, Ni, and Zn) for Enhanced Energy Storage Properties
by Mawuse Amedzo-Adore and Jeong In Han
Nanomaterials 2024, 14(21), 1765; https://doi.org/10.3390/nano14211765 - 3 Nov 2024
Viewed by 984
Abstract
With its distinctive multiple electrochemical reaction, iron vanadate (FeV3O9.2.6H2O) is considered as a promising electrode material for energy storage. However, it has a relatively low practical specific capacitance. Therefore, using the low temperature sol–gel synthesis process, transition [...] Read more.
With its distinctive multiple electrochemical reaction, iron vanadate (FeV3O9.2.6H2O) is considered as a promising electrode material for energy storage. However, it has a relatively low practical specific capacitance. Therefore, using the low temperature sol–gel synthesis process, transition metal doping was used to enhance the electrochemical performance of layered structured FeV3O9.2.6H2O (FVO). According to this study, FVO doped with transition metals with larger interlayer spacing exhibited superior electrochemical performance than undoped FVO. The Mn-doped FVO electrode showed the highest specific capacitance and retention of 143 Fg−1 and 87%, respectively, while the undoped FVO showed 78 Fg−1 and 54%. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) PXRD patterns, (<b>b</b>) Expanded PXRD patterns, (<b>c</b>) FTIR spectra, and (<b>d</b>) N2 adsorption/desorption isotherm of undoped and doped FVO samples compared.</p>
Full article ">Figure 2
<p>FE-SEM and TEM images of (<b>a</b>,<b>f</b>) FVO, (<b>b</b>,<b>g</b>) Co, (<b>c</b>,<b>h</b>) Mn, (<b>d</b>,<b>i</b>) Ni, and (<b>e</b>,<b>j</b>) Zn samples.</p>
Full article ">Figure 3
<p>(<b>a</b>) V 2p spectra, and (<b>b</b>) O 1s spectra of all samples compared.</p>
Full article ">Figure 4
<p>(<b>a</b>) Co 2p spectra of Co-doped FVO; (<b>b</b>) Mn 2p spectra of Mn-doped FVO; (<b>c</b>) Ni 2p spectra of Ni-doped FVO; and (<b>d</b>) Zn 2p spectra of Zn-doped FVO.</p>
Full article ">Figure 5
<p>(<b>a</b>) CV curves at scan rate 2 mVs<sup>−1</sup> in voltage window 0.0–0.4 V; (<b>b</b>) galvanostatic charge–discharge profiles measured at 0.5 Ag<sup>−1</sup> in voltage window 0.0–0.35 V; (<b>c</b>) rate capability; and (<b>d</b>) cycle performances measured at current density 5 Ag<sup>−1</sup>; of all electrodes.</p>
Full article ">Figure 6
<p>(<b>a</b>) Plot of log (v) vs. log (<span class="html-italic">i</span>) (b-value determination), and (<b>b</b>) plot of <span class="html-italic">i/v<sup>1/2</sup></span> vs. <span class="html-italic">v<sup>1/2</sup></span> of all electrodes compared. (<b>c</b>) Contribution ratio of surface-controlled process to the capacitance at various scan rates, and (<b>d</b>) the contribution ratio of the surface-controlled process to the capacitance at a scan rate of 20 mVs<sup>−1</sup> for the Mn-doped FVO electrode.</p>
Full article ">
33 pages, 4683 KiB  
Article
Component Distribution, Shear-Flow Behavior, and Sol–Gel Transition in Mixed Dispersions of Casein Micelles and Serum Proteins
by Hossein Gholamian, Maksym Loginov, Marie-Hélène Famelart, Florence Rousseau, Fabienne Garnier-Lambrouin and Geneviève Gésan-Guiziou
Foods 2024, 13(21), 3480; https://doi.org/10.3390/foods13213480 - 30 Oct 2024
Viewed by 856
Abstract
The shear flow and solid–liquid transition of mixed milk protein dispersions with varying concentrations of casein micelles (CMs) and serum proteins (SPs) are integral to key dairy processing operations, including microfiltration, ultrafiltration, diafiltration, and concentration–evaporation. However, the rheological behavior of these dispersions has [...] Read more.
The shear flow and solid–liquid transition of mixed milk protein dispersions with varying concentrations of casein micelles (CMs) and serum proteins (SPs) are integral to key dairy processing operations, including microfiltration, ultrafiltration, diafiltration, and concentration–evaporation. However, the rheological behavior of these dispersions has not been sufficiently studied. In the present work, dispersions of CMs and SPs with total protein weight fractions (ωPR) of 0.021–0.28 and SP to total protein weight ratios (RSP) of 0.066–0.214 and 1 were prepared by dispersing the respective protein isolates in the permeate from skim milk ultrafiltration and then further concentrated via osmotic compression. The partition of SPs between the CMs and the dispersion medium was assessed by measuring the dry matter content and viscosity of the dispersion medium after separating it from the CMs via ultracentrifugation. The rheological properties were studied at 20 °C via shear rheometry, and the sol–gel transition was characterized via oscillatory measurements. No absorption of SPs by CMs was observed in dispersions with ωPR = 0.083–0.126, regardless of the RSP. For dispersions of SPs with ωPR ≤ 0.21, as well as the dispersion medium of mixed dispersions with ωPR = 0.083–0.126, the high shear- rate-limiting viscosity was described using Lee’s equation with an SP voluminosity (vSP) of 2.09 mL·g−1. For the mixed dispersions with a CM volume fraction of φCM ≤ 0.37, the relative high shear-rate-limiting viscosity was described using Lee’s equation with a CM voluminosity (vCM) of 4.15 mL·g−1 and a vSP of 2.09 mL·g−1, regardless of the RSP. For the mixed dispersions with φCM > 0.55, the relative viscosity increased significantly with an increasing RSP (this was explained by an increase in repulsion between CMs). However, the sol–gel transition was independent of the RSP and was observed at φCM ≈ 0.65. Full article
(This article belongs to the Section Dairy)
Show Figures

Figure 1

Figure 1
<p>Theoretical DM content of the dispersion medium (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) (calculated under the assumption set C (<a href="#foods-13-03480-t002" class="html-table">Table 2</a>) vs. measured total DM content of supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) of mixed milk protein dispersions). Data for suspensions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles). The dashed line corresponds to <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math> = <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>. The determination coefficient for the data fitting by this equation (<span class="html-italic">r</span><sup>2</sup>) is shown near the dashed line. The DM content at the origin is that of PUF (0.0521).</p>
Full article ">Figure 2
<p>(<b>a</b>) Steady-state apparent viscosity (<span class="html-italic">η</span>) as a function of shear rate (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math>) for WPI dispersions (<span class="html-italic">R<sub>SP</sub></span> = 1) prepared by mixing WPI powder with PUF (solid symbols) and following osmotic compression (open symbols); total protein concentrations (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) are shown next to the curves. (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>η</mi> <mo>(</mo> <mrow> <mrow> <mi>γ</mi> </mrow> </mrow> <mo>|</mo> <mo>)</mo> </mrow> </semantics></math> for supernatants produced via ultracentrifugation of mother dispersions; the <span class="html-italic">R<sub>SP</sub></span> in dispersions used for ultracentrifugation are shown next to the curves; the dashed line indicates experimental data for PUF.</p>
Full article ">Figure 3
<p>(<b>a</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of SP concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) in liquid (solid symbols) and compressed (open symbols) dispersions of WPI; solid curve corresponds to data fitting by Lee’s equation, dashed curves correspond to lower and upper limits of the 95% prediction interval; left and bottom axes correspond to results over the entire range of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> studied (circles), while the right and top axes correspond to results for the lowest <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> (diamonds). (<b>b</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of SP concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) in supernatants produced via ultracentrifugation of dispersions with different <span class="html-italic">R<sub>SP</sub></span> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>: <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles); solid and dashed curves are the same as in (<b>a</b>).</p>
Full article ">Figure 4
<p>Steady-state apparent viscosity (<span class="html-italic">η</span>) as a function of shear rate (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math>) for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (<b>a</b>), 0.158 (<b>b</b>), and 0.214 (<b>c</b>): solid symbols—liquid dispersions prepared by mixing and diluting CNI and WPI powders with PUF; open symbols—compressed dispersions prepared via osmotic compression; values of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> are shown next to the curves.</p>
Full article ">Figure 5
<p>Apparent (<span class="html-italic">η</span>) (<b>a</b>) and relative (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>d</mi> <mi>m</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) (<b>b</b>) viscosity of dispersions with near-Newtonian behavior at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> <mo>=</mo> <mn>100</mn> <msup> <mrow> <mi>s</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math> as a function of <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> calculated for <span class="html-italic">v<sub>CM</sub></span> = 4.15 mL·g<sup>−1</sup>: <span class="html-italic">R<sub>SP</sub></span> = 0.066 (squares), 0.158 (triangles), and 0.214 (inverted triangles); the solid curve in <a href="#foods-13-03480-f005" class="html-fig">Figure 5</a>b was calculated using the modified Lee’s equation (Equation (18)).</p>
Full article ">Figure 6
<p>Apparent viscosity (<span class="html-italic">η</span>) measured at <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 10 s<sup>−1</sup> (<b>a</b>) and 100 s<sup>−1</sup> (<b>b</b>) as a function of total protein concentration (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (CNI, squares), 0.158 (triangles), 0.214 (inverted triangles), and 1 (WPI, circles).</p>
Full article ">Figure 7
<p>Relative viscosity of dispersion (<span class="html-italic">η<sub>r</sub></span>) at <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msub> <mover accent="true"> <mrow> <mi>γ</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> of 0.005 Pa (<b>a</b>) and 0.15 Pa (<b>b</b>) as a function of the volume fraction of casein micelles (CMs) in dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 (squares), 0.158 (triangles), and 0.214 (inverted triangles); dashed vertical lines correspond to <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.55 and 0.65.</p>
Full article ">Figure 8
<p>Elastic (<span class="html-italic">G</span>′) (solid symbols) and loss (<span class="html-italic">G</span>″) (open symbols) moduli as a function of frequency (<span class="html-italic">θ</span>) for mixed milk protein dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.006 (CNI) (<b>a</b>), 0.158 (<b>b</b>), 0.214 (<b>c</b>), and 1 (WPI) (<b>d</b>). Symbols correspond to experimental data; total protein concentrations (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math>) are shown next to the curves.</p>
Full article ">Figure 9
<p>Elastic modulus <span class="html-italic">G</span>′ (measured at a frequency (<math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math>) of 0.1 Hz) and exponent <span class="html-italic">m</span> (estimated using Equation (20)) for mixed milk protein dispersions prepared by osmotic compression for <span class="html-italic">R<sub>SP</sub></span> = 0.066 (CNI) (<b>a</b>), 0.158 (<b>b</b>), and 0.214 (<b>c</b>). The left and right edges of shaded zones correspond to dispersions with liquid-like (<span class="html-italic">G</span>′ &lt; <span class="html-italic">G</span>″) or solid-like (<span class="html-italic">G</span>′ &gt; <span class="html-italic">G</span>″) behavior, respectively, in the range of <math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math> of 0.1–10.0 Hz.</p>
Full article ">Figure 10
<p>Composition of dispersions in the sol–gel transition region as a function of the relative concentration of serum proteins (<span class="html-italic">R<sub>SP</sub></span>): particle volume fraction of casein micelles (CMs) in the dispersion (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (squares), the volume fraction of proteins in the dispersion (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (circles), and volume fraction of serum proteins in the dm (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math>) (diamonds). Symbols correspond to the middle of the sol–gel transition region, while error bars indicate its upper and lower limits (shaded zones in <a href="#foods-13-03480-f009" class="html-fig">Figure 9</a>). The dashed horizontal line indicates the mean <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>φ</mi> </mrow> <mrow> <mi>C</mi> <mi>M</mi> </mrow> <mrow> <mi>D</mi> <mo>,</mo> <mi>s</mi> <mi>g</mi> </mrow> </msubsup> </mrow> </semantics></math> (0.65).</p>
Full article ">Figure 11
<p>(<b>a</b>) Theoretical dry matter (DM) content of the dispersion medium (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>t</mi> <mi>h</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) (calculated for the assumed values of <span class="html-italic">K<sub>SP</sub></span>, <span class="html-italic">f<sub>o</sub></span>, and <span class="html-italic">f<sub>eMN</sub></span> listed) vs. measured total DM content of supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>D</mi> <mi>M</mi> <mo>,</mo> <mi>e</mi> <mi>x</mi> <mi>p</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> </mrow> </semantics></math>) of mixed milk protein dispersions. Data for suspensions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>PS</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles). The dashed line is the 1:1 line. The DM content at the origin is that of PUF (0.0521). (<b>b</b>) Relative viscosity (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>r</mi> <mi>P</mi> <mi>U</mi> <mi>F</mi> </mrow> </msub> </mrow> </semantics></math>) at a shear rate of 100 s<sup>−1</sup> as a function of the concentration of serum protein in supernatants (<math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>S</mi> <mi>P</mi> </mrow> <mrow> <mi>d</mi> <mi>m</mi> </mrow> </msubsup> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>) (calculated for the assumed values of <span class="html-italic">K<sub>SP</sub></span>, <span class="html-italic">f<sub>o</sub></span>, and <span class="html-italic">f<sub>eMN</sub></span> listed); data for dispersions with <span class="html-italic">R<sub>SP</sub></span> = 0.066 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.083 and 0.125 (squares), <span class="html-italic">R<sub>SP</sub></span> = 0.158 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.105 and 0.126 (triangles), and <span class="html-italic">R<sub>SP</sub></span> = 0.214 and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>ω</mi> </mrow> <mrow> <mi>P</mi> <mi>R</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msubsup> </mrow> </semantics></math> = 0.126 (inverted triangles); solid curve corresponds to data described by Lee’s equation, while dashed curves correspond to lower and upper limits of the 95% prediction interval.</p>
Full article ">
10 pages, 2857 KiB  
Article
Synthesis and Properties of a Red Na5Zn2Gd1−x(MoO4)6: xEu3+ Phosphor
by Wa Gao, Ren Sha and Jun Ai
Crystals 2024, 14(11), 933; https://doi.org/10.3390/cryst14110933 - 29 Oct 2024
Viewed by 563
Abstract
Novel Eu3+-doped Na5Zn2Gd(MoO4)6 triple molybdate phosphors were fabricated by the sol-gel method. The structure, morphology, and luminescent properties have been characterized by X-ray diffraction (XRD), thermogravimetric differential thermal analysis (TG-DTA), scanning electron microscopy (SEM), [...] Read more.
Novel Eu3+-doped Na5Zn2Gd(MoO4)6 triple molybdate phosphors were fabricated by the sol-gel method. The structure, morphology, and luminescent properties have been characterized by X-ray diffraction (XRD), thermogravimetric differential thermal analysis (TG-DTA), scanning electron microscopy (SEM), FTIR spectroscopy, and luminescence spectroscopy. The results indicated that the synthesized Na5Zn2Gd1−x(MoO4)6: xEu3+ phosphor consisted of a pure phase with monoclinic structure. Under excitation at 465 nm, the Na5Zn2Gd1−x(MoO4)6: xEu3+ phosphor exhibits an intensive red emission band around 610 nm corresponding to the transition of 5D07F2 which is much higher than that 5D07F1 at 594 nm, which was appropriate for a blue LED. According to the influence of the synthesis conditions, the phosphors showed the highest emission intensity when the doping concentration of Eu3+ was 25 mol.% and the molar ratio of citric acid to metal ions was 2:1. Na5Zn2Gd0.75(MoO4)6: 0.25 Eu3+ with the color coordinates (x = 0.658, y = 0.341) is a more stable red phosphor for blue-based white LEDs than the commercial Y2O2S: Eu3+ red phosphor (0.48, 0.50) due to its being closer to the NTSC standard values (0.670, 0.330). Full article
(This article belongs to the Special Issue Optical Properties of Crystalline Semiconductors and Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>TG−DTA curves of the Na<sub>5</sub>Zn<sub>2</sub>Gd(MoO<sub>4</sub>)<sub>6</sub>: Eu<sup>3+</sup>.</p>
Full article ">Figure 2
<p>Structure and morphology of Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>1−x</sub>(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup>: (<b>a</b>) XRD patterns; (<b>b</b>,<b>c</b>) SEM images of Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>0.75</sub>(MoO<sub>4</sub>)<sub>6</sub>: 0.25Eu<sup>3+</sup>; (<b>d</b>) crystal structure.</p>
Full article ">Figure 3
<p>FTIR spectra of the precursor and Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>0.75</sub>(MoO<sub>4</sub>)<sub>6</sub>: 0.25Eu<sup>3+</sup> phosphor calcined at 800 °C.</p>
Full article ">Figure 4
<p>UV–vis diffuse reflectance spectra of Na<sub>5</sub>Zn<sub>2</sub>Gd(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup> (<b>a</b>); Tauc plots (<b>b</b>).</p>
Full article ">Figure 5
<p>The excitation spectrum of Na<sub>5</sub>Zn<sub>2</sub>Gd(MoO<sub>4</sub>)<sub>6</sub>: Eu<sup>3+</sup> sample monitored at λ<sub>em</sub> = 614 nm.</p>
Full article ">Figure 6
<p>The emission spectra of Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>1−x</sub>(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup>.</p>
Full article ">Figure 7
<p>Emission spectra of Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>1−x</sub>(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup> doped with different Eu<sup>3+</sup> concentrations.</p>
Full article ">Figure 8
<p>Emission spectra of different citric acid concentrations in Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>1−x</sub>(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup>.</p>
Full article ">Figure 9
<p>Luminescent decay curves of Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>1−x</sub>(MoO<sub>4</sub>)<sub>6</sub>: xEu<sup>3+</sup> monitored at 615 nm and excited with 465 nm.</p>
Full article ">Figure 10
<p>CIE chromaticity diagram of the Na<sub>5</sub>Zn<sub>2</sub>Gd<sub>0.75</sub>(MoO<sub>4</sub>)<sub>6</sub>: 0.25 Eu<sup>3+</sup>.</p>
Full article ">
11 pages, 9125 KiB  
Article
Improving the Nonvolatile Memory Characteristics of Sol–Gel-Processed Y2O3 RRAM Devices Using Mono-Ethanolamine Additives
by Seongwon Heo, Soohyun Choi, Sangwoo Lee, Yoonjin Cho, Jin-Hyuk Bae, In-Man Kang, Kwangeun Kim, Won-Yong Lee and Jaewon Jang
Materials 2024, 17(21), 5252; https://doi.org/10.3390/ma17215252 - 28 Oct 2024
Viewed by 870
Abstract
In this study, Y2O3-based resistive random-access memory (RRAM) devices with a mono-ethanolamine (MEA) stabilizer fabricated using the sol–gel process on indium tin oxide/glass substrates were investigated. The effects of MEA content on the structural, optical, chemical, and electrical characteristics [...] Read more.
In this study, Y2O3-based resistive random-access memory (RRAM) devices with a mono-ethanolamine (MEA) stabilizer fabricated using the sol–gel process on indium tin oxide/glass substrates were investigated. The effects of MEA content on the structural, optical, chemical, and electrical characteristics were determined. As the MEA content increased, film thickness and crystallite size decreased. In particular, the increase in MEA content slightly decreased the oxygen vacancy concentration. The decreased film thickness decreased the physical distance for conductive filament formation, generating a strong electric field. However, owing to the lowest oxygen vacancy concentration, a large electrical field is required. To ensure data reliability, the endurance cycles across several devices were measured and presented statistically. Additionally, endurance performance improved with the increase in MEA content. Reduced oxygen vacancy concentration can successfully suppress the excess formation of the Ag conductive filament. This simplifies the transition from the high- to the low-resistance state and vice versa, thereby improving the endurance cycles of the RRAM devices. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Cross-sectional SEM image and (<b>b</b>–<b>d</b>) surface SEM images as a function of MEA content.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) Cross-sectional SEM image and (<b>b</b>–<b>d</b>) surface SEM images as a function of MEA content.</p>
Full article ">Figure 2
<p>Three-dimensional SPM surface images of the Y<sub>2</sub>O<sub>3</sub> films, as a function of MEA content: (<b>a</b>) YMEA-0, (<b>b</b>) YMEA-10, and (<b>c</b>) YMEA-20, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) GIXRD data and (<b>b</b>) calculated crystallite sizes of the Y<sub>2</sub>O<sub>3</sub> films as a function of MEA content.</p>
Full article ">Figure 4
<p>(<b>a</b>) Transmittance spectra of the glass/ITO/Y<sub>2</sub>O<sub>3</sub> films and (<b>b</b>) (αhν)<sup>2</sup> versus hν plots of the Y<sub>2</sub>O<sub>3</sub> films as a function of MEA content.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) XPS O1s spectra of the Y<sub>2</sub>O<sub>3</sub> films and (<b>d</b>) chemical composition variations as a function of MEA content.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>–<b>c</b>) XPS O1s spectra of the Y<sub>2</sub>O<sub>3</sub> films and (<b>d</b>) chemical composition variations as a function of MEA content.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic image of the fabricated Y<sub>2</sub>O<sub>3</sub> RRAM devices. (<b>b</b>–<b>d</b>) Representative I–V curves of the Y<sub>2</sub>O<sub>3</sub> RRAM devices as a function of MEA content. The arrows and numbers indicate the voltage sweep directions.</p>
Full article ">Figure 6 Cont.
<p>(<b>a</b>) Schematic image of the fabricated Y<sub>2</sub>O<sub>3</sub> RRAM devices. (<b>b</b>–<b>d</b>) Representative I–V curves of the Y<sub>2</sub>O<sub>3</sub> RRAM devices as a function of MEA content. The arrows and numbers indicate the voltage sweep directions.</p>
Full article ">Figure 7
<p>(<b>a</b>) Obtained performance parameters of the RRAM devices: (<b>a</b>) SET and RESET voltages and (<b>b</b>) HRS and LRS.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>c</b>) Representative endurance data, as a function of MEA content and (<b>d</b>) statistical data of the obtained endurance cycles.</p>
Full article ">Figure 9
<p>Representative retention data as a function of MEA content: (<b>a</b>) YMEA-0, (<b>b</b>) YMEA-10 and (<b>c</b>) YMEA-20, respectively.</p>
Full article ">
29 pages, 11641 KiB  
Article
Process Mapping of the Sol–Gel Transition in Acid-Initiated Sodium Silicate Solutions
by Marzieh Matinfar and John A. Nychka
Gels 2024, 10(10), 673; https://doi.org/10.3390/gels10100673 - 21 Oct 2024
Viewed by 1239
Abstract
Fabricating large-scale porous bioactive glass bone scaffolds presents significant challenges. This study aims to develop formable, in situ setting scaffolds with a practical gelation time of about 10 min by mixing 45S5 bioactive glass with sodium silicate (waterglass) and an acid initiator. The [...] Read more.
Fabricating large-scale porous bioactive glass bone scaffolds presents significant challenges. This study aims to develop formable, in situ setting scaffolds with a practical gelation time of about 10 min by mixing 45S5 bioactive glass with sodium silicate (waterglass) and an acid initiator. The effects of pH (2–11), waterglass concentration (15–50 wt.%), and acid initiator type (phosphoric or boric acid) were examined to optimize gelation kinetics and microstructure. A 10 min gelation time was achieved with boric acid and phosphoric acid at various pH levels by adjusting the waterglass concentration. Exponential and polynomial models were proposed to predict gelation times in basic and acidic environments, respectively. The optical properties of the gels were studied qualitatively and quantitatively, providing insights into gelation kinetics and structure. Acidic gels formed smaller particles in a dense network (pores < 550 nm) with higher light transmittance, while basic gels had larger aggregates (pores ~5 µm) and lower transmittance. As the waterglass concentration decreased, pore size and transmittance converged in both groups. The onset of gelation was detected around 8 min using the derivative of light transmittance. This work identifies the key factors controlling waterglass gelation and their impact on gel structure, enabling the tailored creation of formable, in situ setting bioactive glass bone scaffolds. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Progression from problematization to design approach in this study: (<b>a</b>) three major challenges were spotted applying problematization in bioactive glass bone scaffold processing techniques; (<b>b</b>) to address the problems in section (<b>a</b>), a new processing technique (the previous proof-of-concept work) was developed which resulted in bioactive glass composites bone scaffolds with long setting time impractical for clinal application; (<b>c</b>) developing a new approach using acid-initiated waterglass to reduce the setting time to a practical range for clinical applications.</p>
Full article ">Figure 2
<p>(<b>a</b>) the general trend for gelation time/sol stability versus pH for silica water systems, redrawn from [<a href="#B20-gels-10-00673" class="html-bibr">20</a>]; (<b>b</b>) the general trend for gelation time versus pH for waterglass solution of different concentrations observed in this study; (<b>c</b>) semi-log graph of gelation time versus pH for specimens made of boric acid and phosphoric acid and different initial waterglass concentrations. The target gelation time of 10 min. is marked by the yellow arrow, which intersects the various solutions at different pH values, and hence permits multiple processing routes to fabricate a gel that forms in 10 min. The vertical dashed line presents the point of zero charge (PZC) and isoelectric point (IEP).</p>
Full article ">Figure 3
<p>Dependency of parameters a and b on waterglass concentration. These parameters are part of the fitted equation describing the relationship between gelation time and pH for basic gels (<a href="#gels-10-00673-t001" class="html-table">Table 1</a>). The fitted equations and R-values are shown on the graph.</p>
Full article ">Figure 4
<p>The graph shows the amount of acid initiators (boric acid and phosphoric acid) needed to adjust the pH of silica gels made from waterglass at different concentrations. For clarity, the y-axis is presented in a logarithmic scale.</p>
Full article ">Figure 5
<p>Macrograph of silica gels during sol–gel transition and post-gelation over time. All images were taken at constant white balance and exposure conditions and the blue cast to the images is true to what was usually observed. Gels have varying degrees of transparency, ranging from a clear appearance to a blue cast and opaque.</p>
Full article ">Figure 6
<p>Macrographs showing the Tyndall effect of silica gels during sol–gel transition and post-gelation over time while shining a 550 nm green through the specimens. The laser is causing a reflection above the laser line from the top surface of sol–gel meniscus, or oven the top cap on the vials in some cases. All images were taken at constant white balance and exposure conditions and the green reflection is true to what was usually observed. The changes in light scattering, indicated by the intensity and amount of reflected green light, are inversely correlated with light transmittance in <a href="#gels-10-00673-f005" class="html-fig">Figure 5</a>. This observation suggests that the light not transmitted through the sample is scattering rather than being absorbed.</p>
Full article ">Figure 7
<p>UV/VIS spectra of silica gels during sol–gel transition. The spectra show light transmittance in the visible light region (400–700 nm) at various time intervals after mixing waterglass solutions of different concentrations with boric acid and phosphoric acid. The measurements were taken at the gelation point and up to 60 min after gelation. The figures show decreased light transmittance over time for all gels and less light scattering for acidic gels compared to basic gels.</p>
Full article ">Figure 8
<p>(<b>a</b>) Light transmittance at an average wavelength of 550 nm over time for silica gels made with phosphoric acid and boric acid at different waterglass concentrations (<b>left</b>), light transmittance macrographs, Tyndall effect macrographs obtained by shining a 550 nm laser, and SEM micrographs of the gels, taken 60 min post-gelation (<b>right</b>). Although all gels share the same gelation time of 10 min, they exhibit distinctly different optical properties and microstructures based on processing conditions. (<b>b</b>) The first derivative of light transmittance of silica gels versus time, showing an inflection point around the gelation point. (<b>c</b>) The second derivative of light transmittance of silica gels versus time, used to more clearly identify changes in the light transmittance rate. The maxima at ~8 min indicate the onset of gelation. The dashed line and arrow mark the gelation point at 10 min by tube inversion method in all figures.</p>
Full article ">Figure A1
<p>Water and waterglass solutions before and after CO<sub>2</sub> gassing using a bicycle CO<sub>2</sub> inflator. Each solution contains pH indicator, with the color change representing the decrease in pH due to the introduction of CO<sub>2</sub>.</p>
Full article ">Figure A2
<p>(<b>a</b>) Water solutions containing pH indicator before and after CO<sub>2</sub> gassing using a Soda Stream machine. (<b>b</b>) Waterglass containing pH indicator after CO<sub>2</sub> gassing using a Soda Stream machine, with a layer of gel formed after gassing. (<b>c</b>) The procedure of CO<sub>2</sub> gassing of waterglass–bioactive glass paste inside an open cylindrical mold using a Soda Stream machine. (<b>d</b>) Damage to the waterglass–bioactive glass paste due to high concentrated pressure. (<b>e</b>) We used a long plastic cylinder to increase the distance of the nozzle. However, some specimens were still damaged by the concentrated pressure, as shown in the top right.</p>
Full article ">Figure A3
<p>(<b>a</b>) Illustration detailing the components used to assemble a medical-grade concentrated CO<sub>2</sub> gas system. (<b>b</b>) Image showing a negative silicone mold filled with waterglass, highlighting the formation of a thin silica gel film after exposure to medical-grade CO<sub>2</sub>.</p>
Full article ">Figure A4
<p>(<b>a</b>) The image on the left shows the composite holding its shape after waterglass gelation. The two images on the right demonstrate how the composite deforms under compressive load. (<b>b</b>) These curves compare the mechanical properties of air-set and CO<sub>2</sub>-set waterglass–bioactive glass composites.</p>
Full article ">
15 pages, 1749 KiB  
Review
The Sol–Gel Process, a Green Method Used to Obtain Hybrid Materials Containing Phosphorus and Zirconium
by Petru Merghes, Gheorghe Ilia, Bianca Maranescu, Narcis Varan and Vasile Simulescu
Gels 2024, 10(10), 656; https://doi.org/10.3390/gels10100656 - 13 Oct 2024
Viewed by 1109
Abstract
The sol–gel process is a green method used in the last few decades to synthesize new organic–inorganic phosphorus-containing hybrid materials. The sol–gel synthesis is a green method because it takes place in mild conditions, mostly by using water or alcohol as solvents, at [...] Read more.
The sol–gel process is a green method used in the last few decades to synthesize new organic–inorganic phosphorus-containing hybrid materials. The sol–gel synthesis is a green method because it takes place in mild conditions, mostly by using water or alcohol as solvents, at room temperature. Therefore, the sol–gel method is, among others, a promising route for obtaining metal-phosphonate networks. In addition to phosphorus, the obtained hybrid materials could also contain titanium, zirconium, boron, and other elements, which influence their properties. The sol–gel process has two steps: first, the sol formation, and second, the transition to the gel phase. In other words, the sol–gel process converts the precursors into a colloidal solution (sol), followed by obtaining a network (gel). By using the sol–gel method, different organic moieties could be introduced into an inorganic matrix, resulting in organic–inorganic hybrid structures (sometimes they are also referred as organic–inorganic copolymers). Full article
(This article belongs to the Special Issue Smart Hydrogels: From Rational Design to Applications (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>The nonhydrolytic step and the hydrolytic step take place during the sol–gel process.</p>
Full article ">Figure 2
<p>SEM image of an organic–inorganic hybrid synthesized by the sol–gel method, at a molar ratio of 2:1 (excess of phenyl phosphonic acid) [<a href="#B15-gels-10-00656" class="html-bibr">15</a>].</p>
Full article ">Figure 3
<p>TEM analysis of an organic–inorganic hybrid synthesized by the sol–gel method, at a molar ratio of 2:1 (excess of phenyl phosphonic acid) [<a href="#B15-gels-10-00656" class="html-bibr">15</a>].</p>
Full article ">Figure 4
<p>TEM analysis of an organic–inorganic hybrid synthesized by the sol–gel method, at a 1:1 molar ratio phenyl phosphonic acid: butyl-zirconate [<a href="#B15-gels-10-00656" class="html-bibr">15</a>].</p>
Full article ">Figure 5
<p>Schematic representation of the sol–gel process steps, of condensation and hydrolysis, if an alkyl-zirconate (ZrOR) is used as precursor.</p>
Full article ">
27 pages, 3014 KiB  
Review
Advances in Chitosan-Based Smart Hydrogels for Colorectal Cancer Treatment
by Urszula Piotrowska and Klaudia Orzechowska
Pharmaceuticals 2024, 17(10), 1260; https://doi.org/10.3390/ph17101260 - 25 Sep 2024
Viewed by 1692
Abstract
Despite advancements in early detection and treatment in developed countries, colorectal cancer (CRC) remains the third most common malignancy and the second-leading cause of cancer-related deaths worldwide. Conventional chemotherapy, a key option for CRC treatment, has several drawbacks, including poor selectivity and the [...] Read more.
Despite advancements in early detection and treatment in developed countries, colorectal cancer (CRC) remains the third most common malignancy and the second-leading cause of cancer-related deaths worldwide. Conventional chemotherapy, a key option for CRC treatment, has several drawbacks, including poor selectivity and the development of multiple drug resistance, which often lead to severe side effects. In recent years, the use of polysaccharides as drug delivery systems (DDSs) to enhance drug efficacy has gained significant attention. Among these polysaccharides, chitosan (CS), a linear, mucoadhesive polymer, has shown promise in cancer treatment. This review summarizes current research on the potential applications of CS-based hydrogels as DDSs for CRC treatment, with a particular focus on smart hydrogels. These smart CS-based hydrogel systems are categorized into two main types: stimuli-responsive injectable hydrogels that undergo sol-gel transitions in situ, and single-, dual-, and multi-stimuli-responsive CS-based hydrogels capable of releasing drugs in response to various triggers. The review also discusses the structural characteristics of CS, the methods for preparing CS-based hydrogels, and recent scientific advances in smart CS-based hydrogels for CRC treatment. Full article
(This article belongs to the Special Issue Biodegradable Polymeric Nanosystems for Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of chitin and chitosan. This figure is adapted from [<a href="#B22-pharmaceuticals-17-01260" class="html-bibr">22</a>] with permission under CC BY 4.0 license.</p>
Full article ">Figure 2
<p>The gelation process of CS hydrogel (<b>a</b>) via acidic solvent; (<b>b</b>) via alkaline solvent. This figure is adapted from [<a href="#B79-pharmaceuticals-17-01260" class="html-bibr">79</a>] with permission under CC BY 4.0 license.</p>
Full article ">Figure 3
<p>Preparation of chitosan nanoparticles by ionic gelation method.</p>
Full article ">Figure 4
<p>Unique properties of smart hydrogels.</p>
Full article ">Figure 5
<p>In situ gelation mechanism between chitosan and β-glycerophosphate.</p>
Full article ">Figure 6
<p>Phase transition behavior of HACPN-DOX and HACPN solutions at 25 and 37 °C. This figure is adapted from [<a href="#B119-pharmaceuticals-17-01260" class="html-bibr">119</a>] with permission under CC BY 4.0 license.</p>
Full article ">
24 pages, 18105 KiB  
Article
Diverse Strategies to Develop Poly(ethylene glycol)–Polyester Thermogels for Modulating the Release of Antibodies
by Daria Lipowska-Kur, Łukasz Otulakowski, Urszula Szeluga, Katarzyna Jelonek and Alicja Utrata-Wesołek
Materials 2024, 17(18), 4472; https://doi.org/10.3390/ma17184472 - 12 Sep 2024
Viewed by 975
Abstract
In this work, we present basic research on developing thermogel carriers containing high amounts of model antibody immunoglobulin G (IgG) with potential use as injectable molecules. The quantities of IgG loaded into the gel were varied to evaluate the possibility of tuning the [...] Read more.
In this work, we present basic research on developing thermogel carriers containing high amounts of model antibody immunoglobulin G (IgG) with potential use as injectable molecules. The quantities of IgG loaded into the gel were varied to evaluate the possibility of tuning the dose release. The gel materials were based on blends of thermoresponsive and degradable ABA-type block copolymers composed of poly(lactide-co-glycolide)-b-poly(ethylene glycol)-b-poly(lactide-co-glycolide) (PLGA–PEG–PLGA) or poly(lactide-co-caprolactone)-b-poly(ethylene glycol)-b-(lactide-co-caprolactone) (PLCL–PEG–PLCL). Primarily, the gels with various amounts of IgG were obtained via thermogelation, where the only factor inducing gel formation was the change in temperature. Next, to control the gels’ mechanical properties, degradation rate, and the extent of antibody release, we have tested two approaches. The first one involved the synergistic physical and chemical crosslinking of the copolymers. To achieve this, the hydroxyl groups located at the ends of the PLGA–PEG–PLGA chain were modified into acrylate groups. In this case, the thermogelation was accompanied by chemical crosslinking through the Michael addition reaction. Such an approach increased the dynamic mechanical properties of the gels and simultaneously prolonged their decomposition time. An alternative solution was to suspend crosslinked PEG–polyester nanoparticles loaded with IgG in a PLGA–PEG–PLGA gelling copolymer. We observed that loading IgG into thermogels lowered the gelation temperature (TGEL) value and increased the storage modulus of the gels, as compared with gels without IgG. The prepared gel materials were able to release the IgG from 8 up to 80 days, depending on the gel formulation and on the amount of loaded IgG. The results revealed that additional, chemical crosslinking of the thermogels and also suspension of particles in the polymer matrix substantially extended the duration of IgG release. With proper matching of the gel composition, environmental conditions, and the type and amount of active substances, antibody-containing thermogels can serve as effective IgG delivery materials. Full article
(This article belongs to the Special Issue Applied Stimuli-Responsive Polymer Based Materials)
Show Figures

Figure 1

Figure 1
<p>DMA measurement results for exemplifying <b>P2</b> and <b>P4</b> copolymers at a concentration of 25 wt%: <b>P2</b> in (<b>A</b>) water and (<b>B</b>) 0.9% NaCl; <b>P4</b> in (<b>C</b>) water and (<b>D</b>) 0.9% NaCl (oscillation frequency of 1 Hz).</p>
Full article ">Figure 2
<p>DMA measurement results for exemplifying blends: (<b>A</b>) <b>P2</b>/<b>P4</b> in water, (<b>B</b>) <b>P2</b>/<b>P4</b> in 0.9% NaCl, (<b>C</b>) <b>P2</b>/<b>P4</b> with IgG in 0.9% NaCl, (<b>D</b>) <b>P2</b>/<b>P4</b>/<b>P6</b> in 0.9% NaCl, and (<b>E</b>) <b>P2</b>/<b>P4</b>/<b>P6</b> blend with IgG in 0.9% NaCl (oscillation frequency of 1 Hz).</p>
Full article ">Figure 3
<p>DMA measurement results for (<b>A</b>) <b>P2M</b> in 0.9% NaCl, (<b>B</b>) <b>P4M</b> in 0.9% NaCl (<b>C</b>) <b>P2M</b>/<b>P4M</b> blend in 0.9% NaCl, (<b>D</b>) <b>P2M</b>/<b>P4M</b> blend with IgG in 0.9% NaCl, and (<b>E</b>) <b>P2M</b>/<b>P4M</b>/<b>P6</b> blend in 0.9% NaCl. Oscillation frequency of 1 Hz.</p>
Full article ">Figure 4
<p>(<b>A</b>) Decrease in molar mass during degradation of two-component gels made of non-modified and modified copolymers; (<b>B</b>) comparison of molar mass loss during degradation of <b>P2</b>/<b>P4</b> and <b>P2</b>/<b>P4</b>/<b>P6</b> gels (0.9% NaCl, 33 °C).</p>
Full article ">Figure 5
<p>IgG cumulative release from (<b>A</b>) <b>P2</b>/<b>P4,</b> (<b>B</b>) <b>P2</b>/<b>P4</b>/<b>P6</b>, (<b>C</b>) <b>P2M</b>/<b>P4M</b>, and (<b>D</b>) <b>P2M</b>/<b>P4M</b>/<b>P6</b> blends.</p>
Full article ">Figure 6
<p>CryoTEM images of (<b>A</b>) empty nanoparticles, and representative <b>P2M</b>/<b>P4M</b> nanoparticles loaded with IgG at a 33.6 mg/mL concentration via (<b>B</b>) the diffusion method (IgG_I) and (<b>C</b>) entrapment during formation method (IgG_II).</p>
Full article ">Figure 7
<p>DMA measurement results for (<b>A</b>) <b>P2M</b>/<b>P4M</b> nanoparticles suspended in <b>P3</b> thermogel with IgG (<b>P2M</b>/<b>P4M</b>/<b>P3</b>/IgG_II, 33.6 mg/mL IgG) and (<b>B</b>) blank nanoparticles suspended in <b>P3</b> thermogel. Storage modulus and loss modulus results were obtained at an oscillation frequency of 1 Hz.</p>
Full article ">Figure 8
<p>IgG release plots from the (<b>A</b>) <b>P2M</b>/<b>P4M</b>/<b>P3</b>/IgG_I (<b>B</b>) <b>P2M</b>/<b>P4M</b>/<b>P3</b>/IgG_II system.</p>
Full article ">Figure 9
<p>Effect of <b>P2M</b>/<b>P4M</b>/<b>P6</b> polymer and the extract of the <b>P2M</b>/<b>P4M</b>/<b>P6</b> gel on proliferation of normal human cells after 72 h (<span class="html-italic">p</span> &lt; 0.05 vs. control group).</p>
Full article ">Scheme 1
<p>A schematic representation of the thermogel systems used for release of IgG antibody. Thermogels are composed of a two- or three-component blend of unmodified copolymers (<b>A</b>), unmodified with modified copolymers (<b>B</b>), or of crosslinked nanoparticles embedded in a polymer matrix (<b>C</b>).</p>
Full article ">
13 pages, 7922 KiB  
Article
Strongly Fluorescent Blue-Emitting La2O3: Bi3+ Phosphor for Latent Fingerprint Detection
by Hanen Douiri, Marwa Abid, Lamia Rzouga Haddada, Layla Brini, Alessandra Toncelli, Najoua Essoukri Ben Amara and Ramzi Maalej
Materials 2024, 17(17), 4217; https://doi.org/10.3390/ma17174217 - 26 Aug 2024
Viewed by 855
Abstract
Blue-emitting bismuth-doped lanthanum oxide (La2O3: Bi3+) with various concentrations of Bi was synthesized using the sol–gel combustion method and used for visualization of latent fingerprints (LFPs). An X-ray diffraction (XRD) study revealed the hexagonal structure of the [...] Read more.
Blue-emitting bismuth-doped lanthanum oxide (La2O3: Bi3+) with various concentrations of Bi was synthesized using the sol–gel combustion method and used for visualization of latent fingerprints (LFPs). An X-ray diffraction (XRD) study revealed the hexagonal structure of the phosphors and total incorporation of the bismuth in the La2O3 matrix. Field Emission Scanning Electron Microscopy (FE-SEM) and Fourier Transform Infrared Spectroscopy (FTIR) were used to study the morphology and the relative vibrations of the synthesized samples. Photoluminescence (PL) studies showed strong blue emission around 460 nm due to the 3P11S0 transition. Clear bright-blue fingerprint images were obtained with the powder dusting method on various surfaces like aluminum, compact discs, glass, wood and marble. A first evaluation of these images indicated a clear visualization of all three levels of details and a very high contrast ranging from 0.41 on marble to 0.90 on aluminum. As a further step, we used an algorithm for extracting fingerprint minutiae with which we succeeded in detecting all three levels of fingerprint details and even the most difficult ones, like open and closed pores. According to these analyses, La2O3: Bi phosphor is demonstrated to be an effective blue fluorescent powder for excellent visualization of latent fingerprints. Full article
Show Figures

Figure 1

Figure 1
<p>Description of the preprocessing phase.</p>
Full article ">Figure 2
<p>La<sub>2</sub>O<sub>3</sub> and La<sub>2-x</sub>O<sub>3</sub>: Bi<sub>x</sub> powders (x = 0.002, 0.005 and 0.01).</p>
Full article ">Figure 3
<p>SEM micrographs of prepared La<sub>2</sub>O<sub>3</sub> Bi phosphors: (<b>a</b>,<b>b</b>) for x = 0.002; (<b>c</b>,<b>d</b>) for x = 0.01.</p>
Full article ">Figure 4
<p>FTIR spectra of the La<sub>2</sub>O<sub>3</sub> and La<sub>2-x</sub>O<sub>3</sub>: Bi<sub>x</sub> powders.</p>
Full article ">Figure 5
<p>(a) Excitation and (b) emission spectra of La<sub>2-x</sub>O<sub>3</sub>: Bi<sub>x</sub> doped with different Bi concentrations.</p>
Full article ">Figure 6
<p>CIE color chromaticity coordinates of La<sub>2</sub>O<sub>3</sub>: Bi<sup>3+</sup> phosphors: A (x = 0.2), B (x = 0.5) and C (x = 1).</p>
Full article ">Figure 7
<p>Illustration of the development of latent fingermarks using the 1%Bi: La<sub>2</sub>O<sub>3</sub> phosphor dusting process.</p>
Full article ">Figure 8
<p>Photographs (under 302 nm UV-illumination) of LFPs on different substrates developed by La<sub>2</sub>O<sub>3</sub>: Bi<sup>3+</sup> fluorescent blue powder. For every picture, the line intensity profile along the red line and the best contrast obtained are shown at the bottom.</p>
Full article ">Figure 9
<p>Latent fingerprint image visualization (under 302 nm UV illumination) displaying first-, second- and third-level details.</p>
Full article ">Figure 10
<p>The latent fingerprint image features evidenced by the 1% Bi, La<sub>2</sub>O<sub>3</sub> phosphors on aluminum foil under UV light based on an automatic algorithm ((<b>a</b>) level 1, (<b>b</b>) level 2, and (<b>c</b>) level 3).</p>
Full article ">Figure 10 Cont.
<p>The latent fingerprint image features evidenced by the 1% Bi, La<sub>2</sub>O<sub>3</sub> phosphors on aluminum foil under UV light based on an automatic algorithm ((<b>a</b>) level 1, (<b>b</b>) level 2, and (<b>c</b>) level 3).</p>
Full article ">
Back to TopTop