[go: up one dir, main page]

 
 
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (335)

Search Parameters:
Keywords = XRPD

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 3569 KiB  
Article
A Cippus from Turris Libisonis: Evidence for the Use of Local Materials in Roman Painting on Stone in Northern Sardinia
by Roberta Iannaccone, Stefano Giuliani, Sara Lenzi, Matteo M. N. Franceschini, Silvia Vettori and Barbara Salvadori
Minerals 2024, 14(10), 1040; https://doi.org/10.3390/min14101040 (registering DOI) - 17 Oct 2024
Abstract
The ancient Roman town of Turris Libisonis was located on the northern coast of Sardinia and was known in the past as an important naval port. Located in the Gulf of Asinara, it was a Roman colony from the 1st century BCE and [...] Read more.
The ancient Roman town of Turris Libisonis was located on the northern coast of Sardinia and was known in the past as an important naval port. Located in the Gulf of Asinara, it was a Roman colony from the 1st century BCE and became one of the richest towns on the island. Among the archaeological finds in the area, the cippus exhibited in the Antiquarium Turritano is of great interest for its well-preserved traces of polychromy. The artefact dates back to the early Imperial Age and could have had a funerary or votive function. The artefact was first examined using a portable and non-invasive protocol involving multi-band imaging (MBI), portable X-ray fluorescence (p-XRF), portable FT-IR in external reflectance mode (ER FT-IR) and Raman spectroscopy. After this initial examination, a few microfragments were collected and investigated by optical microscopy (OM), X-ray powder diffraction (XRPD), Fourier-transform infrared spectroscopy in ATR mode (ATR FT-IR) and micro-ATR mode (μATR FT-IR) and Scanning Electron Microscopy/Energy Dispersive Spectroscopy (SEM-EDS) to improve our knowledge and characterize the materials and to determine their provenience. The results contribute to a better understanding of the provenance of materials and shed light on pigments on stone and their use outside the Italian peninsula and, in particular, Roman Sardinia. Full article
(This article belongs to the Special Issue Geomaterials and Cultural Heritage)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The position of the cippus in the Antiquarium of Porto Torres (SS) and the two sides analyzed: (<b>b</b>) side A; (<b>c</b>) side B (courtesy of Ministero della Cultura–Direzione Regionale Musei Sardegna).</p>
Full article ">Figure 2
<p>(<b>a</b>) Raking light detail of flaking areas; (<b>b</b>) High magnification image (60×) of the area in the red square; (<b>c</b>) SEM image in BSE from the sample (red point) and (<b>d</b>) EDS analysis results of a point on recrystallized salt.</p>
Full article ">Figure 3
<p>(<b>a</b>) Raman spectrum of point corresponding to (<b>a</b>) a black area and (<b>b</b>) a yellow area, respectively. On the right, the optical microscopic details of the points (60×). In (<b>b</b>) Raman spectrum of yellow area, in black, and goethite reference spectrum in yellow (RRUFF mineral database).</p>
Full article ">Figure 4
<p>(<b>a</b>) p-XRF spectra of point 8 in black, point 10 in dotted black and the reference background in red; (<b>b</b>) the locations of measured points 8 and 10 are shown (<b>c</b>) Raman spectra of green earth pigment at point 8 (in green) and point 10 (in gray).</p>
Full article ">Figure 5
<p>Microphotographs of thin sections of the carbonate rock of the cippus (<b>a</b>,<b>b</b>) and of the mortar covering the cippus (<b>c</b>,<b>d</b>) (using a polarized light microscope): (<b>a</b>,<b>c</b>) parallel nicols; (<b>b</b>,<b>d</b>) crossed nicols.</p>
Full article ">Figure 6
<p>Geological map of Porto Torres and its surrounding area. Modified from [<a href="#B46-minerals-14-01040" class="html-bibr">46</a>].</p>
Full article ">Figure 7
<p>Macro-photo of carbonate rock of the cippus. Various fossils were observed in the carbonatic rock of the cippus: (<b>a</b>) gastropods; (<b>b</b>) ammonites; (<b>c</b>) algae tallus.</p>
Full article ">Figure 8
<p>Fragment analyzed by SEM−EDS. (<b>a</b>) Optical microscope image at 50×; (<b>b</b>) backscattered details at 127×; and (<b>c</b>) EDS analysis results.</p>
Full article ">Figure 9
<p>(<b>a</b>) FT-IR micro-ATR spectrum obtained from the green sample using point analysis with a TE-MCT detector and (<b>b</b>) FT-IR FPA-ATR spectrum extracted from the chemical map shown above. The arrow indicates the position of the spectrum.</p>
Full article ">
21 pages, 4283 KiB  
Article
Beta-Hydroxybutyric Acid as a Template for the X-ray Powder Diffraction Analysis of Gamma-Hydroxybutyric Acid
by Domenica Marabello, Carlo Canepa, Alma Cioci and Paola Benzi
Molecules 2024, 29(19), 4678; https://doi.org/10.3390/molecules29194678 - 2 Oct 2024
Viewed by 390
Abstract
In this paper, we report the possibility of using the X-ray powder diffraction (XRPD) technique to detect gamma-hydroxybutyric acid (GHB) in the form of its sodium salt in different beverages, but because it is not possible to freely buy GHB, beta-hydroxybutyric acid (BHB) [...] Read more.
In this paper, we report the possibility of using the X-ray powder diffraction (XRPD) technique to detect gamma-hydroxybutyric acid (GHB) in the form of its sodium salt in different beverages, but because it is not possible to freely buy GHB, beta-hydroxybutyric acid (BHB) and its sodium salt (NaBHB) were used as a model to fine-tune an X-ray diffraction method for the qualitative analysis of the sodium salt of GHB. The method requires only a small quantity of beverage and an easy sample preparation that consists only of the addition of NaOH to the drink and a subsequent drying step. The dry residue obtained can be easily analyzed with XRPD using a single-crystal X-ray diffractometer, which exploits its high sensitivity and allows for very fast pattern collection. Several beverages with different NaBHB:NaOH molar ratios were tested, and the results showed that NaBHB was detected in all drinks analyzed when the NaBHB:NaOH molar ratio was 1:50, using a characteristic peak at very low 2θ values, which also permitted the detection of its presence in complex beverage matrices. Moreover, depending on the amount of NaOH added, shifting and/or splitting of the characteristic NaBHB salt peak was observed, and the origin of this behavior was investigated. Full article
(This article belongs to the Section Analytical Chemistry)
Show Figures

Figure 1

Figure 1
<p>The XRD pattern of NaBHB obtained from BHB + NaOH in an aqueous solution and NaBHB and NaOH recrystallized from aqueous solutions. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 2
<p>The XRPD patterns of the solids obtained from different beverages spiked with 20% <span class="html-italic">v</span>/<span class="html-italic">v</span> of BHB and a 1:1 BHB:NaOH ratio (series A). The pattern of NaBHB obtained from the BHB aqueous solution is also reported for comparison. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 3
<p>The XRPD patterns of the residues of some beverages spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> of BHB and a 1:2 BHB:NaOH ratio (series B1). The pattern of the standard NaBHB obtained from the BHB aqueous solution is also reported for comparison. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 4
<p>The X-ray patterns of the solids obtained from some beverages spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> BHB and a BHB:NaOH ratio of 1:5 (series C), 1:10 (series D), 1:20 (series E), and 1:50 (series F). The patterns of NaBHB obtained from the BHB aqueous solution and of recrystallized NaOH are also shown. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 4 Cont.
<p>The X-ray patterns of the solids obtained from some beverages spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> BHB and a BHB:NaOH ratio of 1:5 (series C), 1:10 (series D), 1:20 (series E), and 1:50 (series F). The patterns of NaBHB obtained from the BHB aqueous solution and of recrystallized NaOH are also shown. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 5
<p>The X-ray patterns of dry residues from the scotch samples spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> BHB and BHB:NaOH ratios 1:5, 1:10, 1:20, and 1:50. The inset highlights the region of 2θ values between 5° and 20°. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 6
<p>The X-ray patterns of the dry residues obtained from water solutions spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> BHB and BHB:NaOH ratios from 1:2 to 1:50. The inset highlights the region of 2θ values between 5° and 20°. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 7
<p>The XRPD patterns of some non-alcoholic (<b>A</b>), alcoholic (<b>B</b>), and cocktail beverages (<b>C</b>). The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure 7 Cont.
<p>The XRPD patterns of some non-alcoholic (<b>A</b>), alcoholic (<b>B</b>), and cocktail beverages (<b>C</b>). The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure A1
<p>The XRPD patterns of the solids obtained from different beverages spiked with 20% <span class="html-italic">v</span>/<span class="html-italic">v</span> NaBHB and a 1:1 NaBHB:NaOH molar ratio. The pattern of NaBHB recrystallized from an aqueous solution is also reported for comparison.</p>
Full article ">Figure A2
<p>The XRPD patterns of the residues of some beverages spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> NaBHB and a 1:2 NaBHB:NaOH molar ratio. The pattern of NaBHB recrystallized from an aqueous solution is also reported for comparison. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">Figure A3
<p>The XRPD patterns of the residues of some beverages without BHB but spiked with the same NaOH content of series B samples. The pattern of NaOH recrystallized from aqueous solution is also re-ported for comparison. The patterns have been translated on the ordinate axis.</p>
Full article ">Figure A4
<p>The X-ray patterns of dry residues from the rum samples spiked with 1% <span class="html-italic">v</span>/<span class="html-italic">v</span> BHB and BHB:NaOH molar ratios of 1:5, 1:10, 1:20, and 1:50. The inset highlights the region between 5° and 20° 2θ values. The patterns have been translated on the ordinate axis and normalized with respect to the higher signal of NaBHB.</p>
Full article ">
12 pages, 14597 KiB  
Article
Influences of La2O3 Addition on Connectivity of Phase Compositions and Microstructural Evolution of Weld Slag
by Xiaoyu He, Min Zhang, Longyu Lei and Yi Li
Crystals 2024, 14(10), 841; https://doi.org/10.3390/cryst14100841 - 27 Sep 2024
Viewed by 335
Abstract
In this work, the influences of La2O3 addition on the connectivity of the glass network, phase compositions and microstructural evolution of weld slag were investigated through Raman spectrum, X-ray powder diffraction (XRPD), SEM and EBSD technologies. All experimental results indicated [...] Read more.
In this work, the influences of La2O3 addition on the connectivity of the glass network, phase compositions and microstructural evolution of weld slag were investigated through Raman spectrum, X-ray powder diffraction (XRPD), SEM and EBSD technologies. All experimental results indicated that La2O3 addition could modify the whole glass network’s connectivity and short-ordered units. According to the Raman spectrum, only 1 wt.% La2O3 addition resulted in the occurrence of a unique linking mode of Si-O and Al-O tetrahedrons that was assigned to feldspar phases (albite or anorthite). Further XRPD examination showed that the primary phases were albite and anorthite, which agreed with the Raman results. Moreover, enhanced linkage between Si-O and Al-O tetrahedrons needs a large amount Na+ to achieve electric neutrality. This repaired the connectivity of the slag network due to the lack of Na+. Additionally, the solubility of La2O3 in the slag matrix was limited to about 3 wt.%. Adding further La2O3 to this weld slag, the existence form of La2O3 retained its original status. Thus, La2O3 addition that exceeded 3 wt.% had little effect on the slag connectivity besides providing a phase interface. In conclusion, La2O3 addition in weld slag could promote the hardness and the formation of feldspar phases. Feldspar is an extremely fragile silicon aluminate crystal. The factors mentioned above caused the detachability to be enhanced when adding La2O3. Full article
(This article belongs to the Section Materials for Energy Applications)
Show Figures

Figure 1

Figure 1
<p>DSC curves of welding slag with La<sub>2</sub>O<sub>3</sub> addition.</p>
Full article ">Figure 2
<p>Raman profiles. (<b>A</b>) Raman profiles of welding slag with La<sub>2</sub>O<sub>3</sub> addition; (<b>B</b>) residual curves of Raman profiles of welding slag with La<sub>2</sub>O<sub>3</sub> addition.</p>
Full article ">Figure 3
<p>DOP values of welding slag with La<sub>2</sub>O<sub>3</sub> addition.</p>
Full article ">Figure 4
<p>XRD patterns of welding slag with La<sub>2</sub>O<sub>3</sub> addition.</p>
Full article ">Figure 5
<p>Patterns of welding slag with La<sub>2</sub>O<sub>3</sub> added under 40 h crystallization treatment.</p>
Full article ">Figure 6
<p>Microstructure of welding slag with La<sub>2</sub>O<sub>3</sub> addition: (<b>A</b>) 0 wt.%; (<b>B</b>) 1 wt.%; (<b>C</b>) 3 wt.%; (<b>D</b>) 5 wt.%.</p>
Full article ">Figure 7
<p>Phase distribution of welding slag after 40 h crystallization treatment with 5 wt.% La<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 8
<p>Phase ID of welding slag with 5 wt.% La<sub>2</sub>O<sub>3</sub> added under 40 h crystallization treatment. (<b>A</b>) SEM image, (<b>B</b>) EDS results of selected area (marked with red cross in (<b>A</b>)), (<b>C</b>) Raw Kikuchi band, (<b>D</b>) Indexed Kikuchi band.</p>
Full article ">
23 pages, 4779 KiB  
Article
An Additive Manufacturing MicroFactory: Overcoming Brittle Material Failure and Improving Product Performance through Tablet Micro-Structure Control for an Immediate Release Dose Form
by Elke Prasad, John Robertson and Gavin W. Halbert
Polymers 2024, 16(18), 2566; https://doi.org/10.3390/polym16182566 - 11 Sep 2024
Viewed by 604
Abstract
Additive manufacturing of pharmaceutical formulations offers advanced micro-structure control of oral solid dose (OSD) forms targeting not only customised dosing of an active pharmaceutical ingredient (API) but also custom-made drug release profiles. Traditionally, material extrusion 3D printing manufacturing was performed in a two-step [...] Read more.
Additive manufacturing of pharmaceutical formulations offers advanced micro-structure control of oral solid dose (OSD) forms targeting not only customised dosing of an active pharmaceutical ingredient (API) but also custom-made drug release profiles. Traditionally, material extrusion 3D printing manufacturing was performed in a two-step manufacturing process via an intermediate feedstock filament. This process was often limited in the material space due to unsuitable (brittle) material properties, which required additional time to develop complex formulations to overcome. The objective of this study was to develop an additive manufacturing MicroFactory process to produce an immediate release (IR) OSD form containing 250 mg of mefenamic acid (MFA) with consistent drug release. In this study, we present a single-step additive manufacturing process employing a novel, filament-free melt extrusion 3D printer, the MicroFactory, to successfully print a previously ‘non-printable’ brittle Soluplus®-based formulation of MFA, resulting in targeted IR dissolution profiles. The physico-chemical properties of 3D printed MFA-Soluplus®-D-sorbitol formulation was characterised by thermal analysis, Fourier Transform Infrared spectroscopy (FTIR), and X-ray Diffraction Powder (XRPD) analysis, confirming the crystalline state of mefenamic acid as polymorphic form I. Oscillatory temperature and frequency rheology sweeps were related to the processability of the formulation in the MicroFactory. 3D printed, micro-structure controlled, OSDs showed good uniformity of mass and content and exhibited an IR profile with good consistency. Fitting a mathematical model to the dissolution data correlated rate parameters and release exponents with tablet porosity. This study illustrates how additive manufacturing via melt extrusion using this MicroFactory not only streamlines the manufacturing process (one-step vs. two-step) but also enables the processing of (brittle) pharmaceutical immediate-release polymers/polymer formulations, improving and facilitating targeted in vitro drug dissolution profiles. Full article
(This article belongs to the Special Issue Applications of 3D Printing for Polymers, 3rd Edition)
Show Figures

Figure 1

Figure 1
<p>Filament performance in 3D printer drive gear and hot end. (<b>A</b>) Filament conveyed by drive gear. Failure modes in drive gear: (<b>B</b>) brittle failure, (<b>C</b>) ductile failure/buckling, and (<b>D</b>) soft filament shearing. Hot end: (<b>E</b>) material conveyed in hot end; (<b>F</b>) buckling in hot end of 3D printer.</p>
Full article ">Figure 2
<p>Impact of filament dimensions on heat transfer in hot end and dosing accuracy: (<b>A</b>) low diameter, poor heat transfer, and under dosing; (<b>B</b>) target diameter, good heat transfer, and accurate dosing; and (<b>C</b>) diameter in excess of hot end diameter.</p>
Full article ">Figure 3
<p>Successive prints of tablet micro-structure A (47.3% infill, no top or bottom layer) with 50MFA at 140 °C, layer height of 0.2 mm and 0.4 mm nozzle.</p>
Full article ">Figure 4
<p>Micro-structures of 3D printed tablets containing 50% <span class="html-italic">w</span>/<span class="html-italic">w</span> MFA (22 mm × 12 mm × 5 mm): (<b>A</b>) infill 47.3%, no top, no bottom layer; (<b>B</b>) infill 40.6%, no top layer; and (<b>C</b>) infill 37.5%, top and bottom layers. Scale bar: 5 mm.</p>
Full article ">Figure 5
<p>HME-3DP process data. (<b>A</b>) Maximum die pressure (bar) (open diamonds) and (<b>B</b>) maximum torque (%) (open circle) versus number of prints.</p>
Full article ">Figure 5 Cont.
<p>HME-3DP process data. (<b>A</b>) Maximum die pressure (bar) (open diamonds) and (<b>B</b>) maximum torque (%) (open circle) versus number of prints.</p>
Full article ">Figure 6
<p>Oscillatory temperature sweep of 50MFA-SOL. (<b>A</b>) Complex viscosity and (<b>B</b>) storage (G′) and loss (G″) modulus vs. temperature. Storage modulus: filled; loss modulus: open. 3D printed discs (140 °C) shown in blue, 50MFA extrudate processed at 125 °C [<a href="#B7-polymers-16-02566" class="html-bibr">7</a>] shown in black.</p>
Full article ">Figure 7
<p>Thermal analysis of 3D printed disc. Glass transition midpoint (Tg, °C), exothermic peak (°C), and endothermic peak (°C) values for first (light bars) and second (dark bars) heating cycle (n = 2).</p>
Full article ">Figure 8
<p>Thermogram of 3D printed discs. Top dashed lines: first heating cycle from 0 to 250 °C; middle solid line: cooling cycle from 250 to 0 °C; and bottom dash-dotted line: second heating cycle from 0 to 250 °C (all at rate of 20 °C/min) (n = 2).</p>
Full article ">Figure 9
<p>XRPD patterns for (a) 3D printed tablet containing 50% <span class="html-italic">w</span>/<span class="html-italic">w</span> MFA, (b) MFA form III [<a href="#B56-polymers-16-02566" class="html-bibr">56</a>], (c) MFA form II [<a href="#B57-polymers-16-02566" class="html-bibr">57</a>], and (d) MFA form I [<a href="#B58-polymers-16-02566" class="html-bibr">58</a>].</p>
Full article ">Figure 10
<p>Drug release (%) of (<b>A</b>) 50MFA 3D printed tablets (n = 6) over time with different tablet micro-structures: ‘tablet A’—infill 47.3%, no top or bottom layer (grey triangle); ‘tablet B’—infill 40.6%, no top layer (green circle); and ‘tablet C’—infill 37.5%, top and bottom layer (blue square). (<b>B</b>) Powder-filled capsule with 250 mg MFA (n = 6) (Pharmvit Limited (PVL), Batch 4348) [<a href="#B18-polymers-16-02566" class="html-bibr">18</a>]. Red dashed line: 85% MFA released.</p>
Full article ">Figure 11
<p>(<b>A</b>) Tablet porosity and (<b>B</b>) surface area-to-volume ratio (SA/V) vs. model parameters (release exponent, <span class="html-italic">n</span>; shape parameter, <span class="html-italic">kd</span>).</p>
Full article ">
27 pages, 6045 KiB  
Article
Nanostructured Molecular–Network Arsenoselenides from the Border of a Glass-Forming Region: A Disproportionality Analysis Using Complementary Characterization Probes
by Oleh Shpotyuk, Malgorzata Hyla, Adam Ingram, Yaroslav Shpotyuk, Vitaliy Boyko, Pavlo Demchenko, Renata Wojnarowska-Nowak, Zdenka Lukáčová Bujňáková and Peter Baláž
Molecules 2024, 29(16), 3948; https://doi.org/10.3390/molecules29163948 - 21 Aug 2024
Viewed by 651
Abstract
Binary AsxSe100−x alloys from the border of a glass-forming region (65 < x < 70) subjected to nanomilling in dry and dry–wet modes are characterized by the XRPD, micro-Raman scattering (micro-RS) and revised positron annihilation lifetime (PAL) methods complemented by [...] Read more.
Binary AsxSe100−x alloys from the border of a glass-forming region (65 < x < 70) subjected to nanomilling in dry and dry–wet modes are characterized by the XRPD, micro-Raman scattering (micro-RS) and revised positron annihilation lifetime (PAL) methods complemented by a disproportionality analysis using the quantum–chemical cluster modeling approach. These alloys are examined with respect to tetra-arsenic biselenide As4Se2 stoichiometry, realized in glassy g-As65Se35, glassy–crystalline g/c-As67Se33 and glassy–crystalline g/c-As70Se30. From the XRPD results, the number of rhombohedral As and cubic arsenolite As2O3 phases in As-Se alloys increases after nanomilling, especially in the wet mode realized in a PVP water solution. Nanomilling-driven amorphization and reamorphization transformations in these alloys are identified by an analysis of diffuse peak halos in their XRPD patterning, showing the interplay between the levels of a medium-range structure (disruption of the intermediate-range ordering at the cost of an extended-range one). From the micro-RS spectroscopy results, these alloys are stabilized by molecular thioarsenides As4Sen (n = 3, 4), regardless of their phase composition, remnants of thioarsenide molecules destructed under nanomilling being reincorporated into a glass network undergoing a polyamorphic transition. From the PAL spectroscopy results, volumetric changes in the wet-milled alloys with respect to the dry-milled ones are identified as resulting from a direct conversion of the bound positron–electron (Ps, positronium) states in the positron traps. Ps-hosting holes in the PVP medium appear instead of positron traps, with ~0.36–0.38 ns lifetimes ascribed to multivacancies in the As-Se matrix. The superposition of PAL spectrum peaks and tails for pelletized PVP, unmilled, dry-milled, and dry–wet-milled As-Se samples shows a spectacular smoothly decaying trend. The microstructure scenarios of the spontaneous (under quenching) and activated (under nanomilling) decomposition of principal network clusters in As4Se2-bearing arsenoselenides are recognized. Over-constrained As6·(2/3) ring-like network clusters acting as pre-cursors of the rhombohedral As phase are the main products of this decomposition. Two spontaneous processes for creating thioarsenides with crystalline counterparts explain the location of the glass-forming border in an As-Se system near the As4Se2 composition, while an activated decomposition process for creating layered As2Se3 structures is responsible for the nanomilling-driven molecular-to-network transition. Full article
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The normalized XRPD patterns of MQ-derived g/c-As<sub>67</sub>Se<sub>33</sub> in unmilled and dry-milled state showing three principle diffuse peak halos responsible for the FSDP (~15–25°2<span class="html-italic">θ</span>), SSDP (~28–33°2<span class="html-italic">θ</span>) and TDP (~50–60°2<span class="html-italic">θ</span>). The Bragg-diffraction reflexes of crystalline counterparts are reproduced (from the top to the bottom) in a sequence: rhombohedral (grey) As (JCPDS No. 72–1048) [<a href="#B30-molecules-29-03948" class="html-bibr">30</a>,<a href="#B31-molecules-29-03948" class="html-bibr">31</a>], orthorhombic As<sub>4</sub>Se<sub>3</sub> (JCPDS No. 04–4979) [<a href="#B32-molecules-29-03948" class="html-bibr">32</a>], monoclinic As<sub>4</sub>Se<sub>4</sub> (JCPDS No. 71–0388) [<a href="#B33-molecules-29-03948" class="html-bibr">33</a>,<a href="#B34-molecules-29-03948" class="html-bibr">34</a>], monoclinic As<sub>2</sub>Se<sub>3</sub> (JCPDS No. 65–2365) [<a href="#B35-molecules-29-03948" class="html-bibr">35</a>,<a href="#B36-molecules-29-03948" class="html-bibr">36</a>] and trigonal Se (JCPDS No. 73–0465) [<a href="#B37-molecules-29-03948" class="html-bibr">37</a>].</p>
Full article ">Figure 2
<p>The normalized XRPD patterns of MQ-derived g/c-As<sub>67</sub>Se<sub>33</sub> before (<b>a</b>) and after nanomilling in dry (<b>b</b>) and combined dry–wet (<b>c</b>) mode showing three principal diffuse peak halos in comparison with the Bragg-diffraction reflexes from rhombohedral As (JCPDS No. 72–1048) [<a href="#B30-molecules-29-03948" class="html-bibr">30</a>,<a href="#B31-molecules-29-03948" class="html-bibr">31</a>] and cubic arsenolite As<sub>2</sub>O<sub>3</sub> phase (JCPDS No. 36–1490) [<a href="#B38-molecules-29-03948" class="html-bibr">38</a>].</p>
Full article ">Figure 3
<p>The normalized XRPD patterns of g-As<sub>65</sub>Se<sub>35</sub> (black curve) and g/c-As<sub>70</sub>Se<sub>30</sub> (red curve) after nanomilling in dry–wet mode, showing three principal diffuse peak halos corresponding to ‘amorphous’ phase overlapped with sharp broadened Bragg-diffraction reflexes originated from the planes (111) at 13.86°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 6.390Å, <span class="html-italic">I</span> = 63%), (222) at 27.90°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 3.195Å, <span class="html-italic">I</span> = 100%), (400) at 32.33°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 2.769Å, <span class="html-italic">I</span> = 27%), (331) at 35.32°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 2.541 Å, <span class="html-italic">I</span> = 38%), (440) at 46.36°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 1.957Å, <span class="html-italic">I</span> = 27%) and (551) at 59.59°2<span class="html-italic">θ</span> (<span class="html-italic">d</span> = 1.551 Å, <span class="html-italic">I</span> = 27%) in cubic structure of arsenolite As<sub>2</sub>O<sub>3</sub> (JCPDS No. 36–1490) [<a href="#B38-molecules-29-03948" class="html-bibr">38</a>].</p>
Full article ">Figure 4
<p>The normalized micro-RS spectra of MQ-derived g/c-As<sub>67</sub>Se<sub>33</sub> reproduced in a sequence from the bottom to the top: (<b>a</b>) unmilled bulk pieces (unpelletized); (<b>b</b>) pelletized coarse-grained sample; (<b>c</b>) pelletized dry-milled sample; (<b>d</b>) pelletized dry–wet-milled sample. The most prominent features in the micro-RS spectrum of the bulk unmilled sample (<b>a</b>) are marked by vertical arrows, and traced by dotted lines to the respective micro-RS spectra of the pelletized samples (<b>b</b>–<b>d</b>).</p>
Full article ">Figure 5
<p>The normalized micro-RS spectra of MQ-derived g-As<sub>65</sub>Se<sub>35</sub> in unmilled state (<b>a</b>) and after nanomilling in a single dry mode (<b>b</b>). The most prominent features in the micro-RS spectrum of the bulk sample (<b>a</b>) are marked by arrows and traced by dotted lines to that of dry-milled sample (<b>b</b>).</p>
Full article ">Figure 6
<p>The raw PAL spectra of pelletized g/c-As<sub>70</sub>Se<sub>30</sub> in unmilled state (<b>a</b>) and after nanomilling in dry mode (<b>b</b>) and combined dry–wet mode (<b>c</b>) as compared with the spectrum of PVP pelletized under the same conditions (<b>d</b>). The collected PAL spectra are reconstructed from unconstrained three-term fitting and reproduced at background of source contribution with bottom insets showing statistical scatter of variance. The occupation of “tail” states in unmilled and dry-milled samples grows notably under transition to dry–wet-milled sample approaching that in the pelletized PVP.</p>
Full article ">Figure 7
<p>The overlapping of the PAL spectra in the examined arsenoselenides g-As<sub>65</sub>Se<sub>35</sub> (<b>a</b>), g/c-As<sub>67</sub>Se<sub>33</sub> (<b>b</b>) and g/c-As<sub>70</sub>Se<sub>30</sub> (<b>c</b>) pelletized before nanomilling (black points) and after nanomilling in dry mode (red points) and dry–wet mode (green points) as compared with the PAL spectrum in the PVP sample pelletized under the same conditions (blue points). The insets show a nearly invariant tendency in the PAL spectra peaks depressed in the right wing after nanomilling in dry–wet mode due to moderated Ps-formation probability and slightly changed average positron lifetime <span class="html-italic">τ<sub>av</sub></span>. The changes in the PAL spectra tails of unmilled, dry- and dry–wet-milled samples are due to increase in density of o-Ps hosting holes. There is no evident empty gap between the PAL spectra tails for dry-milled and dry–wet-milled glassy samples as compared with glassy-crystalline samples caused by changes in Ps decaying states under transition to annihilation in PVP-bearing medium.</p>
Full article ">Figure 8
<p>The ball-and-stick presentation of optimized configuration of tetra-arsenic biselenide thioarsenide As<sub>4</sub>Se<sub>2</sub>-I molecule composed by four (As-As) bonds in <span class="html-italic">zig-zag</span> sequence (<b>a</b>) and As<sub>4</sub>Se<sub>2</sub>-II molecule composed by (As-As) bond attached to As<sub>3</sub> triangle (<b>b</b>), as compared with As<sub>4</sub>Se<sub>3</sub>H<sub>2</sub> and As<sub>4</sub>Se<sub>4</sub>H<sub>4</sub> molecular prototypes of network clusters derived from these molecules by single (x1-As<sub>4</sub>Se<sub>2</sub>-I—(<b>c</b>), x1-As<sub>4</sub>Se<sub>2</sub>-II—(<b>d</b>)) and double (x2-As<sub>4</sub>Se<sub>2</sub>-I—(<b>e</b>), x2-As<sub>4</sub>Se<sub>2</sub>-II—(<b>f</b>)) breaking in available Se atom positions. The cluster-forming energies <span class="html-italic">E<sub>f</sub></span> are given in respect to AsSe<sub>3/2</sub> pyramid (<span class="html-italic">E<sub>f</sub></span> = −72.309 kcal/mol [<a href="#B60-molecules-29-03948" class="html-bibr">60</a>]). The H, Se and As atoms are, respectively, grey-, blue- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks. The average number of constraints <span class="html-italic">n<sub>c</sub></span> is given following the Phillips-Thorpe constraint-counting algorithm [<a href="#B65-molecules-29-03948" class="html-bibr">65</a>,<a href="#B66-molecules-29-03948" class="html-bibr">66</a>,<a href="#B67-molecules-29-03948" class="html-bibr">67</a>].</p>
Full article ">Figure 9
<p>The ball-and-stick presentation of optimized configuration of tetra-arsenic monoselenide As<sub>4</sub>Se molecule composed by two edge-sharing As<sub>3</sub> triangles (<b>a</b>), as compared with As<sub>4</sub>Se<sub>2</sub>H<sub>2</sub> molecular prototype of network-forming cluster derived from this molecule by breaking in Se atom position x1-As<sub>4</sub>Se (<b>b</b>). The terminated H atoms are grey-colored, Se and As atoms are respectively blue- and red-colored, chemical bonds between atoms are denoted by, respectively, colored sticks.</p>
Full article ">Figure 10
<p>The ball-and-stick presentation of optimized configuration of regular pyramid-shaped tetra-arsenic As<sub>4</sub> thioarsenide molecule (<b>a</b>) and As<sub>6</sub>H<sub>6</sub> molecular prototype of network-forming cluster derived by distortion from this molecule in a form of flattened pyramidal-shaped unit, which composes two-dimensional double-layer network of chair-configurated six-fold rings of As atoms (As<sub>6⋅(2/3)</sub> = As<sub>4</sub>). The H and As atoms are respectively grey- and red-colored, and chemical bonds between atoms are denoted by respectively colored sticks.</p>
Full article ">Figure 11
<p>Three decomposition scenarios of x2-As<sub>4</sub>Se<sub>2</sub>-I network-forming clusters governing molecular-network disproportionality in tetra-arsenic biselenide As<sub>4</sub>Se<sub>2</sub>-bearing arsenoselenides: (<b>a</b>)—spontaneous decomposition under Δ<span class="html-italic">E<sub>f</sub></span> = –0.43 kcal/mol; (<b>b</b>)—spontaneous decomposition under Δ<span class="html-italic">E<sub>f</sub> </span>= –0.17 kcal/mol; (<b>c</b>)—activated decomposition under Δ<span class="html-italic">E<sub>f</sub></span> = +0.375 kcal/mol. The terminated H atoms are grey-colored, Se and As atoms are, respectively, blue- and red-colored, and chemical bonds between atoms are denoted by the respective colored sticks.</p>
Full article ">Figure 12
<p>The unimodal particle size distribution in nanosuspension of MQ-derived g-As<sub>65</sub>Se<sub>35</sub> showing the parameters (x50) and (x99), respectively, approaching ~182 nm and ~291 nm.</p>
Full article ">
18 pages, 2619 KiB  
Article
Development, Characterization, and Cellular Toxicity Evaluation of Solid Dispersion-Loaded Hydrogel Based on Indomethacin
by Zaid Dahma, Alexandra Ibáñez-Escribano, Cristina Fonseca-Berzal, Juan José García-Rodríguez, Covadonga Álvarez-Álvarez, Carlos Torrado-Salmerón, Santiago Torrado-Santiago and Paloma Marina de la Torre-Iglesias
Polymers 2024, 16(15), 2174; https://doi.org/10.3390/polym16152174 - 30 Jul 2024
Viewed by 628
Abstract
Indomethacin (IND) as a non-selective cyclooxygenase 1 and 2 inhibitor administered orally causes numerous adverse effects, mostly related to the gastrointestinal tract. Moreover, when applied exogenously in topical preparations, there are obstacles to its permeation through the stratum corneum due to its low [...] Read more.
Indomethacin (IND) as a non-selective cyclooxygenase 1 and 2 inhibitor administered orally causes numerous adverse effects, mostly related to the gastrointestinal tract. Moreover, when applied exogenously in topical preparations, there are obstacles to its permeation through the stratum corneum due to its low water solubility and susceptibility to photodegradation. In this work, solid dispersions (SDs) of IND with low-substituted hydroxypropyl cellulose (LHPC) were developed. The IND—SDs were incorporated into a hydroxypropyl guar (HPG) hydrogel to enhance drug solubility on the skin. The hydrogels were characterized by scanning electron microscopy (SEM), differential scanning calorimetry (DSC), powder X-ray diffraction (XRPD), Fourier-transform infrared spectroscopy (FTIR), viscosity, drug release, and unspecific cytotoxicity in mammalian cells. SEM showed a highly porous structure for SD hydrogels. DSC and XRPD studies showed that amorphous IND species were formed; therefore, these hydrogels exhibited superior drug release in comparison with IND raw material hydrogels. FTIR evidenced the presence of the hydrogen bond in the SD hydrogel. The rheology parameter viscosity increased across gels formulated with SDs in comparison with hydrogels with pure IND. In addition, IND—SD hydrogels combine the advantages of a suitable viscosity for dermal use and no potentially hazardous skin irritation. This study suggests that the formulated IND—SD hydrogels represent a suitable candidate for topical administration. Full article
(This article belongs to the Special Issue Advanced Biopolymer-Based Composites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scanning electron micrographs of the following samples: (<b>A</b>) IND—RM; (<b>B</b>) PM—1:2.5; (<b>C</b>) SD—1:1; (<b>D</b>) SD—0:2.5; (<b>E</b>) SD—1:2.5; (<b>F</b>) hydrogel composed of pure IND (HIND—RM); (<b>G</b>) physical mixture hydrogel HPM—1:2.5; and (<b>H</b>) solid dispersion hydrogel HSD—1:2.5.</p>
Full article ">Figure 2
<p>DSC thermograms of IND and LHPC raw materials and physical mixture hydrogel (HPM—1:2.5), HPG hydrogel (H—HPG), and solid dispersion hydrogels (HSD—1:0, HSD—1:1, HSD—1:2.5, HSD—0:2.5, HSD—1:5, and HSD—1:10).</p>
Full article ">Figure 3
<p>X-ray powder diffraction scans of IND and LHPC raw materials, physical mixture hydrogel (HPM—1:2.5), HPG hydrogel (H—HPG), and solid dispersion hydrogels (HSD—1:0, HSD—1:1, HSD—1:2.5, HSD—0:2.5, HSD—1:5, and HSD—1:10).</p>
Full article ">Figure 4
<p>FTIR spectra of the pure IND (IND—RM), HPG hydrogel (H—HPG), hydrogel composed of pure IND (HIND—RM), physical mixture hydrogel HPM—1:2.5, and solid dispersion hydrogel HSD—1:2.5.</p>
Full article ">Figure 5
<p>Viscosity behavior of hydrogel formulations: HPG blank hydrogel (H—HPG), hydrogel composed of pure IND (HIND—RM), physical mixture hydrogel HPM—1:2.5, and solid dispersion hydrogel HSD—1:2.5. Results are presented as mean values (n = 3) for each formulation.</p>
Full article ">Figure 6
<p>Solubility studies were conducted on IND raw material (IND—RM), solid dispersions at ratios of IND:LHPC 1:0 (SD—1:0), 1:1 (SD—1:1), 1:2.5 (SD—1:2.5), 1:5 (SD—1:5), and 1:10 (SD—1:10), as well as the physical mixture with a IND:LHPC ratio of 1:2.5 (PM—1:2.5) in phosphate buffer (pH 5.8). ANOVA was used for multiple comparison. (*) indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05) compared with IND—RM. (#) indicates a significant difference (<span class="html-italic">p</span> &lt; 0.05) compared with SD—1:1.</p>
Full article ">Figure 7
<p>(<b>A</b>) Release profiles of IND at pH 5.8 for hydrogels composed of pure IND (HIND—RM), physical mixture hydrogel (HPM—1:2.5), and solid dispersion hydrogels (HSD—1:0, HSD—1:1, HSD—1:2.5, HSD—1:5, and HSD—1:10); (<b>B</b>) histogram of drug release (%) at 120 min for different hydrogel formulations (mean ± standard deviation, n = 4). ANOVA was used for multiple comparison. (*) indicates significant difference (<span class="html-italic">p</span> &lt; 0.05) compared with HIND—RM, and (#) indicates significant difference (<span class="html-italic">p</span> &lt; 0.05) compared with HSD—1:0.</p>
Full article ">
20 pages, 3446 KiB  
Article
Molecular-Network Transformations in Tetra-Arsenic Triselenide Glassy Alloys Tuned within Nanomilling Platform
by Oleh Shpotyuk, Malgorzata Hyla, Yaroslav Shpotyuk, Zdenka Lukáčová Bujňáková, Peter Baláž, Pavlo Demchenko, Andrzej Kozdraś, Vitaliy Boyko and Andriy Kovalskiy
Molecules 2024, 29(14), 3245; https://doi.org/10.3390/molecules29143245 - 9 Jul 2024
Cited by 1 | Viewed by 748
Abstract
Polyamorphic transformations driven by high-energy mechanical ball milling (nanomilling) are recognized in a melt-quenched glassy alloy of tetra-arsenic triselenide (As4Se3). We employed XRPD analysis complemented by thermophysical heat-transfer and micro-Raman spectroscopy studies. A straightforward interpretation of the medium-range structural [...] Read more.
Polyamorphic transformations driven by high-energy mechanical ball milling (nanomilling) are recognized in a melt-quenched glassy alloy of tetra-arsenic triselenide (As4Se3). We employed XRPD analysis complemented by thermophysical heat-transfer and micro-Raman spectroscopy studies. A straightforward interpretation of the medium-range structural response to milling-driven reamorphization is developed within a modified microcrystalline model by treating diffuse peak-halos in the XRPD patterns of this alloy as a superposition of the Bragg-diffraction contribution from inter-planar correlations, which are supplemented by the Ehrenfest-diffraction contribution from inter-atomic and/or inter-molecular correlations related to derivatives of thioarsenide As4Sen molecules, mainly dimorphite-type As4Se3 ones. These cage molecules are merely destroyed under milling, facilitating the formation of a polymerized network with enhanced calorimetric heat-transfer responses. Disruption of intermediate-range ordering, due to weakening of the FSDP (the first sharp diffraction peak), accompanied by an enhancement of extended-range ordering, due to fragmentation of structural entities responsible for the SSDP (the second sharp diffraction peak), occurs as an interplay between medium-range structural levels in the reamorphized As4Se3 glass alloy. Nanomilling-driven destruction of thioarsenide As4Sen molecules followed by incorporation of their remnants into a glassy network is proved by micro-Raman spectroscopy. Microstructure scenarios of the molecular-to-network polyamorphic transformations caused by the decomposition of the As4Se3 molecules and their direct destruction under grinding are recognized by an ab initio quantum-chemical cluster-modeling algorithm. Full article
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The XRPD patterns of unmilled MQ-derived and dry-milled g-As<sub>57</sub>Se<sub>43</sub> alloys showing regions of the three most prominent diffuse peak-halos corresponding to the FSDP (~15–25°2<span class="html-italic">θ</span>), SSDP (~28–33°2<span class="html-italic">θ</span>) and TDP (~50–60°2<span class="html-italic">θ</span>). Theoretical Bragg-diffraction reflexes of monoclinic As<sub>2</sub>Se<sub>3</sub> (JCPDS No. 65-2365) [<a href="#B27-molecules-29-03245" class="html-bibr">27</a>,<a href="#B28-molecules-29-03245" class="html-bibr">28</a>], monoclinic As<sub>4</sub>Se<sub>4</sub> (JCPDS No. 71-0388), orthorhombic As<sub>4</sub>Se<sub>3</sub> (JCPDS No. 04-4979), trigonal Se (JCPDS No. 73-0465) and rhombohedral As (JCPDS No. 72-1048) are reproduced below for comparison (see text for more details).</p>
Full article ">Figure 2
<p>The reconstructed structural fragment of orthorhombic As<sub>4</sub>Se<sub>3</sub> showing (<b>a</b>)—arrangement of cage-like As<sub>4</sub>Se<sub>3</sub> molecules in respect to the family of (111) planes corresponding to the strongest Bragg-diffraction line, (<b>b</b>)—As<sub>4</sub>Se<sub>3</sub> molecule (centered in B) in surrounding of 12 neighbors forming B[B<sub>12</sub>] anticubooctahedron; (<b>c</b>)—possible inter-molecular centroid-centroid distances B-B within B[B<sub>12</sub>] polyhedron (in Å). The averaged B-B distance around each ‘dummy atom’ derived from hexagonal close packing of 12 molecules d<sub>B-B</sub>(As<sub>4</sub>Se<sub>3</sub>) approaches 6.650 Ǻ (see text for more details).</p>
Full article ">Figure 3
<p>The hf-DSC thermograms detected in the dynamic heating-run regime show the variations in non-reversing DSC heat flow <span class="html-italic">HF</span><sub>nrev</sub> in unmilled (black curve) and nanomilled (red curve) As<sub>4</sub>Se<sub>3</sub>. The glass transition temperature <span class="html-italic">T</span><sub>g</sub> enhancement in this molecular-network glassy alloy undergoing nanomilling-driven amorphous-I-to-amorphous-II (reamorphization) transition is obvious.</p>
Full article ">Figure 4
<p>The micro-RS spectra collected from unmilled (black curve) and nanomilled (red curve) samples of g-As<sub>4</sub>Se<sub>3</sub> (the most prominent RS-active bands in unmilled glass alloy are distinguished).</p>
Full article ">Figure 5
<p>The optimized ball-and-stick presentation of three types of tetra-arsenic triselenide (As<sub>4</sub>Se<sub>3</sub>) cage molecules possessing different arrangement of neighboring As atoms: (<b>a</b>)—dimorphite-type As<sub>4</sub>Se<sub>3</sub>-I in <span class="html-italic">triangular-pyramidal</span> (As<sub>3</sub>)-As configuration; (<b>b</b>)—As<sub>4</sub>Se<sub>3</sub>-II in <span class="html-italic">chain</span>-like (<span class="html-italic">zig-zag</span>) As4 configuration; (<b>c</b>)—As<sub>4</sub>Se<sub>3</sub>-III in <span class="html-italic">star</span>-like As(As)3 configuration. The Se and As atoms are blue- and red-colored, and bonds between these atoms are denoted by respective colored sticks. The cluster-forming energies <span class="html-italic">E<sub>f</sub></span> are determined with respect to the energy of a single AsSe<sub>3/2</sub> pyramid.</p>
Full article ">Figure 6
<p>The ball-and-stick presentations of main products of decomposition reaction in g-As<sub>4</sub>Se<sub>3</sub>: (<b>a</b>)—realgar-type As<sub>4</sub>Se<sub>4</sub> molecule possessing cross-orthogonal arrangement of two (As-As) bonds; (<b>b</b>)—H-saturated molecular prototype of network cluster derived from As<sub>4</sub>Se<sub>2</sub> molecule by double breaking in Se atom positions conserving closed <span class="html-italic">tetragon</span>-like As<sub>4</sub> arrangement of (As-As) bonds (As<sub>4</sub>Se<sub>4</sub>H<sub>4</sub>). The terminated H atoms are grey-colored, Se and As atoms are, respectively, blue- and red-colored, and covalent bonds between atoms are denoted by respective colored sticks. The cluster-forming energies <span class="html-italic">E<sub>f</sub></span> are defined with respect to the energy of a single AsSe<sub>3/2</sub> pyramid.</p>
Full article ">Figure 7
<p>Ball-and-stick presentations of decomposition reaction changing molecular-network disproportionality in glassy arsenoselenides compositionally approaching tetra-arsenic triselenide. The two most favorable dimorphite-type As<sub>4</sub>Se<sub>3</sub>-I molecules are transformed into realgar-type As<sub>4</sub>Se<sub>4</sub> molecules and the network-forming remainder becomes closer to amorphous a-As<sub>4</sub>Se<sub>2</sub>. The optimized configurations of molecular and network clusters are reproduced with Se and As atoms, respectively, labeled by blue- and red-colored balls, and terminated H atoms are labeled by grey balls. The decomposition barrier Δ<span class="html-italic">E<sub>f</sub></span> derived from respective cluster-forming energies tends to ~0.41 kcal/mol.</p>
Full article ">Figure 8
<p>The optimized ball-and-stick presentation of most favorable molecular prototypes derived from As<sub>4</sub>Se<sub>3</sub> molecules by breaking in Se positions, which conserve triangular-pyramidal (As<sub>3</sub>)-As (<b>a</b>); chain-like As<sub>4</sub> (<b>b</b>) and star-like As(As)<sub>3</sub> (<b>c</b>) configurations. The terminated H atoms are grey-colored, Se and As atoms are blue- and red-colored, and covalent bonds between atoms are denoted by respective colored sticks. The cluster-forming energies <span class="html-italic">E<sub>f</sub></span> are defined with respect to the energy of a single trigonal AsSe<sub>3/2</sub> pyramid.</p>
Full article ">Figure 9
<p>Positioning of experimental (red points) and calculated (black line) XRPD profiles in the MQ-derived g-As<sub>57</sub>Se<sub>43</sub> showing diffuse peak-halos arrangement with respect to the characteristic inter-planar and inter-atomic (inter-molecular) correlations from quasi-crystalline arsenoselenide remnants (the difference is depicted by the blue curve at the bottom). The insert shows the comparison of “amorphous” halos and most prominent “crystalline” peaks in vitreous As<sub>2</sub>S<sub>3</sub>, As<sub>2</sub>Se<sub>3</sub> and As<sub>2</sub>Te<sub>3</sub> (from the <span class="html-italic">top</span> to the <span class="html-italic">bottom</span>) modified from the known work of Vaipolin and Porai-Koshits, 1963 [<a href="#B53-molecules-29-03245" class="html-bibr">53</a>].</p>
Full article ">
18 pages, 14824 KiB  
Article
Stability Studies of Amorphous Ibrutinib Prepared Using the Quench-Cooling Method and Its Dispersions with Soluplus®
by Igor Mucha, Bożena Karolewicz and Agata Górniak
Polymers 2024, 16(14), 1961; https://doi.org/10.3390/polym16141961 - 9 Jul 2024
Cited by 1 | Viewed by 905
Abstract
The successful development of an amorphous form of a drug demands the use of process conditions and materials that reduce their thermodynamic instability. For the first time, we have prepared amorphous ibrutinib using the quench-cooling method with very high process efficiency. In the [...] Read more.
The successful development of an amorphous form of a drug demands the use of process conditions and materials that reduce their thermodynamic instability. For the first time, we have prepared amorphous ibrutinib using the quench-cooling method with very high process efficiency. In the presented study, different formulations of amorphous active pharmaceutical ingredient (API) with Soluplus (SOL) in various weight ratios 1:9, 3:7, and 1:1 were prepared. The obtained samples were stored under long-term (25 ± 2 °C/60%RH ± 5% RH, 12 months) and accelerated (40 ± 2 °C/75%RH ± 5% RH, 6 months) storage conditions. The physical stability of amorphous ibrutinib and ibrutinib–Soluplus formulations was analyzed using differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), powder X-ray diffraction analysis (XRPD), Fourier transform infrared spectroscopy (FTIR), and scanning electron microscopy (SEM). The lack of significant interactions between the ingredients of the formulation was confirmed by FTIR analysis. An increase in moisture content with an increasing SOL weight ratio was observed under accelerated aging and long-term conditions. Additionally, a slight increase in the moisture content of the stored sample compared to that at the initial time was observed. The results revealed the physical strength of the polymeric systems in the presence of high humidity and temperature. The observed high thermal stability allows the use of various technological processes without the risk of thermal degradation. Full article
(This article belongs to the Special Issue Polymers and Their Role in Drug Delivery)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of ibrutinib (<b>a</b>) and Soluplus<sup>®</sup> (<b>b</b>).</p>
Full article ">Figure 2
<p>Thermal analysis of initial materials: (<b>a</b>) TG curve of IBR raw, (<b>b</b>) DSC curve of IBR raw, (<b>c</b>) TG curve of IBR_0, (<b>d</b>) DSC curve of IBR_0, (<b>e</b>) TG curve of SOL_0, and (<b>f</b>) DSC curve of SOL_0.</p>
Full article ">Figure 3
<p>TG curves of the investigated formulations at initial time: (<b>a</b>) IBR:SOL 1:1_0, (<b>b</b>) IBR:SOL 3:7_0, and (<b>c</b>) IBR:SOL 1:9_0.</p>
Full article ">Figure 4
<p>TG curves of samples stored for 6 months under accelerated conditions: (<b>a</b>) IBR_acc, (<b>b</b>) IBR:SOL 1:1_acc, (<b>c</b>) IBR:SOL 3:7_acc, and (<b>d</b>) IBR:SOL 1:9_acc.</p>
Full article ">Figure 5
<p>TG curves of samples stored for 12 months under long-term conditions: (<b>a</b>) IBR_long, (<b>b</b>) IBR:SOL 1:1_long, (<b>c</b>) IBR:SOL 3:7_long, and (<b>d</b>) IBR:SOL 1:9_long.</p>
Full article ">Figure 6
<p>XRPD patterns of raw and amorphous IBR, Soluplus<sup>®</sup>, and IBR:SOL formulations at initial time and after stability tests.</p>
Full article ">Figure 7
<p>FTIR spectra of raw and amorphous IBR, Soluplus<sup>®</sup>, and IBR:SOL formulations at initial time and after stability tests.</p>
Full article ">Figure 8
<p>SEM images of (<b>a</b>) raw IBR and (<b>b</b>) Soluplus<sup>®</sup>, at initial time.</p>
Full article ">Figure 9
<p>Ibrutinib SEM images of (<b>a</b>) amorphous IBR at initial time, (<b>b</b>) amorphous IBR stored for 6 months, (<b>c</b>) amorphous IBR stored for 12 months, (<b>d</b>) IBR:SOL 1:1 at initial time, (<b>e</b>) IBR:SOL 1:1 stored 6 months, (<b>f</b>) IBR:SOL 1:1 stored for 12 months, (<b>g</b>) IBR:SOL 3:7 at initial time, (<b>h</b>) IBR:SOL 3:7 stored for 6 months (<b>i</b>) IBR:SOL 3:7 stored for 12 months, (<b>j</b>) IBR:SOL 1:9 at initial time, (<b>k</b>) IBR:SOL 1:9 stored for 6 months, and (<b>l</b>) IBR:SOL 1:9 stored for 12 months.</p>
Full article ">Figure 9 Cont.
<p>Ibrutinib SEM images of (<b>a</b>) amorphous IBR at initial time, (<b>b</b>) amorphous IBR stored for 6 months, (<b>c</b>) amorphous IBR stored for 12 months, (<b>d</b>) IBR:SOL 1:1 at initial time, (<b>e</b>) IBR:SOL 1:1 stored 6 months, (<b>f</b>) IBR:SOL 1:1 stored for 12 months, (<b>g</b>) IBR:SOL 3:7 at initial time, (<b>h</b>) IBR:SOL 3:7 stored for 6 months (<b>i</b>) IBR:SOL 3:7 stored for 12 months, (<b>j</b>) IBR:SOL 1:9 at initial time, (<b>k</b>) IBR:SOL 1:9 stored for 6 months, and (<b>l</b>) IBR:SOL 1:9 stored for 12 months.</p>
Full article ">Figure 9 Cont.
<p>Ibrutinib SEM images of (<b>a</b>) amorphous IBR at initial time, (<b>b</b>) amorphous IBR stored for 6 months, (<b>c</b>) amorphous IBR stored for 12 months, (<b>d</b>) IBR:SOL 1:1 at initial time, (<b>e</b>) IBR:SOL 1:1 stored 6 months, (<b>f</b>) IBR:SOL 1:1 stored for 12 months, (<b>g</b>) IBR:SOL 3:7 at initial time, (<b>h</b>) IBR:SOL 3:7 stored for 6 months (<b>i</b>) IBR:SOL 3:7 stored for 12 months, (<b>j</b>) IBR:SOL 1:9 at initial time, (<b>k</b>) IBR:SOL 1:9 stored for 6 months, and (<b>l</b>) IBR:SOL 1:9 stored for 12 months.</p>
Full article ">
13 pages, 8448 KiB  
Article
Effect of Silver Vanadate Nanowires Addition on Structural and Morphological Properties of Dental Porcelain Prepared from Economic Raw Materials
by Badr Eddine Sakhkhane, Marieta Mureșan-Pop, Lucian Barbu-Tudoran, Tamás Lovász and Liliana Bizo
Crystals 2024, 14(7), 616; https://doi.org/10.3390/cryst14070616 - 3 Jul 2024
Viewed by 960
Abstract
In addition to many materials, silver vanadate (AgVO3) has gained interest due to its antimicrobial properties, which opens up the potential for use as an antibacterial agent for biomedical applications. This work aimed to study the effect of AgVO3 addition [...] Read more.
In addition to many materials, silver vanadate (AgVO3) has gained interest due to its antimicrobial properties, which opens up the potential for use as an antibacterial agent for biomedical applications. This work aimed to study the effect of AgVO3 addition on the structural and morphological properties of a developed dental porcelain (DP) prepared from natural raw materials. AgVO3 nanowires, prepared by the coprecipitation method, were added in different amounts (1, 3, and 5 wt.%) to a DP mass with the initial composition of 80 wt.% feldspar, 15 wt.% quartz, and 5 wt.% kaolin, obtained by sintering the mixture at 1300 °C. The structural and morphological properties of AgVO3 and DP were investigated by X-ray powder diffraction (XRPD), Fourier transform infrared spectroscopy (FTIR), scanning electron microscopy/energy-dispersive X-ray spectroscopy (SEM/EDS), and transmission electron microscopy (TEM). The results showed the formation of α-AgVO3 nanowires coated with semispherical metallic silver nanoparticles. Moreover, α-AgVO3 additions do not influence the structural and morphological properties of DP, with the presence of Ag and V clearly identified in the DP with the α-AgVO3 addition. Our findings highlight the potential of this novel material for use in various dental applications. Future studies need to establish the antibacterial properties of the prepared dental material. Full article
(This article belongs to the Special Issue Ceramics: Processes, Microstructures, and Properties)
Show Figures

Figure 1

Figure 1
<p>XRPD powder patterns of the α-AgVO<sub>3</sub> sample, experimental and simulated from COD database.</p>
Full article ">Figure 2
<p>FTIR spectrum of the α-AgVO<sub>3</sub> nanowires.</p>
Full article ">Figure 3
<p>SEM images of α-AgVO<sub>3</sub> at different magnifications: (<b>a</b>) ×1.00 k, (<b>b</b>) ×5.00 k, (<b>c</b>) ×15.0 k, and (<b>d</b>) ×50.0 k, and (<b>e</b>) diameter distribution histogram of nanowires from image (<b>b</b>).</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image at ×100 LM magnification, (<b>b</b>) corresponding EDS spectrum, and (<b>c</b>–<b>f</b>) SEM/EDS elemental mapping of the AgVO<sub>3</sub> sample. In the elemental mapping images, the assignment of color for each element is the following: yellow for Ag, red for V, and blue for O, respectively.</p>
Full article ">Figure 5
<p>TEM images of AgVO<sub>3</sub> at (<b>a</b>) ×350 k and (<b>b</b>) ×40.0 k magnifications.</p>
Full article ">Figure 6
<p>XRPD patterns of DP with (<b>a</b>) 0, (<b>b</b>) 1, (<b>c</b>) 3, and (<b>d</b>) 5 wt% α-AgVO<sub>3</sub> addition, sintered at 1300 °C (Q-quartz).</p>
Full article ">Figure 7
<p>FTIR spectra of DP with (<b>a</b>) 0, (<b>b</b>) 1, (<b>c</b>) 3, and (<b>d</b>) 5 wt.% of α-AgVO<sub>3</sub> addition, sintered at 1300 °C.</p>
Full article ">Figure 8
<p>SEM (<b>left</b>) and EDX (<b>right</b>) analysis realized for the DP samples, with (<b>a</b>) 0, (<b>b</b>) 1, (<b>c</b>) 3, and (<b>d</b>) 5 wt.% addition of α-AgVO<sub>3.</sub> The surface of each composition was analyzed by EDX with circles indicating the chosen area.</p>
Full article ">
20 pages, 12050 KiB  
Article
High C-Rate Performant Electrospun LiFePO4/Carbon Nanofiber Self-Standing Cathodes for Lithium-Ion Batteries
by Debora Maria Conti, Claudia Urru, Giovanna Bruni, Pietro Galinetto, Benedetta Albini, Vittorio Berbenni and Doretta Capsoni
Electrochem 2024, 5(2), 223-242; https://doi.org/10.3390/electrochem5020014 - 5 Jun 2024
Viewed by 1255
Abstract
In the present study, LiFePO4/CNF self-standing cathodes for LIBs are synthesized by electrospinning. A lower active material amount (12.3 and 34.5 wt%) is used, compared to the conventional tape-casted cathodes (70–85 wt%). The characterization techniques (XRPD, SEM, TEM, EDS, Raman spectroscopy, [...] Read more.
In the present study, LiFePO4/CNF self-standing cathodes for LIBs are synthesized by electrospinning. A lower active material amount (12.3 and 34.5 wt%) is used, compared to the conventional tape-casted cathodes (70–85 wt%). The characterization techniques (XRPD, SEM, TEM, EDS, Raman spectroscopy, and thermogravimetry) confirm that the olivine-type structure of LiFePO4 is maintained in the binder-free electrodes, and the active material is homogeneously dispersed into and within the carbon nanofibers. The electrochemical investigation demonstrates that higher Li+ diffusion coefficients (1.36 × 10−11 cm2/s) and improved reversibility are reached for free-standing electrodes, compared to the LiFePO4 tape-casted cathode (80 wt% of active material) appositely prepared for comparison. The 34.5 wt% LiFePO4 self-standing cathode displays a lower capacity fading, good reversibility and stability, enhanced capacity values at C-rates higher than 5C, and a good lifespan when cycled 1000 cycles at 1C and further cycled up to 20C, compared to the tape-casted counterpart. Notably, the improved electrochemical performances are obtained by only the 34.5 wt% of active material. The results evidence the relevant role of the CNF matrix suitable to host LiFePO4, to promote electrolyte permeation and contact with the active material, and to increase the electronic conductivity. Full article
(This article belongs to the Collection Feature Papers in Electrochemistry)
Show Figures

Figure 1

Figure 1
<p>Scheme of the synthetic approaches used for the electrodes’ preparation.</p>
Full article ">Figure 2
<p>X-ray patterns of LiFePO<sub>4</sub>, 10%LiFePO<sub>4</sub>/CNF, 30%LiFePO<sub>4</sub>/CNF, and pure CNF carbonized at 750 °C.</p>
Full article ">Figure 3
<p>Raman spectra of LiFePO<sub>4</sub>, 10%LiFePO<sub>4</sub>/CNF, 30%LiFePO<sub>4</sub>/CNF, and pure CNF, carbonized at 750 °C.</p>
Full article ">Figure 4
<p>SEM ((<b>a</b>): 30kX) and TEM (<b>b</b>) images of LiFePO<sub>4</sub>.</p>
Full article ">Figure 5
<p>SEM images of the surface ((<b>a</b>): 9kX, (<b>b</b>): 25kX) and cross-section ((<b>c</b>): 130X) of the 10%LiFePO<sub>4</sub>/CNF sample. SEM images of the surface ((<b>d</b>): 9kX, (<b>e</b>): 25kX) and cross-section ((<b>f</b>): 170X) of the 30%LiFePO<sub>4</sub>/CNF sample.</p>
Full article ">Figure 6
<p>TEM images of the 10%LiFePO<sub>4</sub>/CNF (<b>a</b>,<b>b</b>) and 30%LiFePO<sub>4</sub>/CNF (<b>c</b>,<b>d</b>) samples.</p>
Full article ">Figure 7
<p>EDS analysis. 10%LiFePO<sub>4</sub>/CNF sample: investigated surface (<b>a</b>) and Fe (<b>b</b>) and P (<b>c</b>) distribution maps; 30%LiFePO<sub>4</sub>/CNF sample: investigated surface (<b>d</b>) and Fe (<b>e</b>) and P (<b>f</b>) distribution maps.</p>
Full article ">Figure 8
<p>EDS analysis. 10%LiFePO<sub>4</sub>/CNF sample: investigated cross-section (<b>a</b>) and Fe (<b>b</b>) and P (<b>c</b>) distribution maps; 30%LiFePO<sub>4</sub>/CNF sample: investigated cross-section (<b>d</b>) and Fe (<b>e</b>) and P (<b>f</b>) distribution maps. The white frame in (<b>a</b>) and (<b>d</b>) indicates the mapped portion.</p>
Full article ">Figure 9
<p>TGA curves of 10%LiFePO<sub>4</sub>/CNF (red), 30%LiFePO<sub>4</sub>/CNF (green), and LiFePO<sub>4</sub> (blue).</p>
Full article ">Figure 10
<p>Cyclic voltammetry and charge/discharge curve of LiFePO<sub>4</sub> (<b>a</b>,<b>b</b>), 10%LiFePO<sub>4</sub>/CNF (<b>c</b>,<b>d</b>), and 30%LiFePO<sub>4</sub>/CNF (<b>e</b>,<b>f</b>).</p>
Full article ">Figure 11
<p>CV and capacitive/diffusive contributions at different scan rates of LiFePO<sub>4</sub> (<b>a</b>,<b>b</b>), 10%LiFePO<sub>4</sub>/CNF (<b>c</b>,<b>d</b>), and 30%LiFePO<sub>4</sub>/CNF (<b>e</b>,<b>f</b>) samples.</p>
Full article ">Figure 12
<p>Charge/discharge cycles at different C-rates of LiFePO<sub>4</sub> (<b>a</b>), 10%LiFePO<sub>4</sub>/CNF (<b>b</b>), 30%LiFePO<sub>4</sub>/CNF (<b>c</b>), and comparisons of the electrochemical performance (<b>d</b>).</p>
Full article ">Figure 13
<p>Long charge/discharge cycles of LiFePO<sub>4</sub> tape-casted (<b>a</b>) and 30%LiFePO<sub>4</sub>/CNF self-standing (<b>b</b>) cathodes.</p>
Full article ">
18 pages, 8116 KiB  
Article
Structural Evolution of the Pharmaceutical Peptide Octreotide upon Controlled Relative Humidity and Temperature Variation
by Maria Athanasiadou, Christina Papaefthymiou, Angelos Kontarinis, Maria Spiliopoulou, Dimitrios Koutoulas, Marios Konstantopoulos, Stamatina Kafetzi, Kleomenis Barlos, Kostas K. Barlos, Natalia Dadivanyan, Detlef Beckers, Thomas Degen, Andrew N. Fitch and Irene Margiolaki
SynBio 2024, 2(2), 205-222; https://doi.org/10.3390/synbio2020012 - 4 Jun 2024
Cited by 1 | Viewed by 1020
Abstract
Octreotide is the first synthetic peptide hormone, consisting of eight amino acids, that mimics the activity of somatostatin, a natural hormone in the body. During the past decades, advanced instrumentation and crystallographic software have established X-Ray Powder Diffraction (XRPD) as a valuable tool [...] Read more.
Octreotide is the first synthetic peptide hormone, consisting of eight amino acids, that mimics the activity of somatostatin, a natural hormone in the body. During the past decades, advanced instrumentation and crystallographic software have established X-Ray Powder Diffraction (XRPD) as a valuable tool for extracting structural information from biological macromolecules. The latter was demonstrated by the successful structural determination of octreotide at a remarkably high d-spacing resolution (1.87 Å) (PDB code: 6vc1). This study focuses on the response of octreotide to different humidity levels and temperatures, with a particular focus on the stability of the polycrystalline sample. XRPD measurements were accomplished employing an Anton Paar MHC-trans humidity-temperature chamber installed within a laboratory X’Pert Pro diffractometer (Malvern Panalytical). The chamber is employed to control and maintain precise humidity and temperature levels of samples during XRPD data collection. Pawley analysis of the collected data sets revealed that the octreotide polycrystalline sample is remarkably stable, and no structural transitions were observed. The compound retains its orthorhombic symmetry (space group: P212121, a = 18.57744(4) Å, b = 30.17338(6) Å, c = 39.70590(9) Å, d ~ 2.35 Å). However, a characteristic structural evolution in terms of lattice parameters and volume of the unit cell is reported mainly upon controlled relative humidity variation. In addition, an improvement in the signal-to-noise ratio in the XRPD data under a cycle of dehydration/rehydration is reported. These results underline the importance of considering the impact of environmental factors, such as humidity and temperature, in the context of structure-based drug design, thereby contributing to the development of more effective and stable pharmaceutical products. Full article
Show Figures

Figure 1

Figure 1
<p>Pawley fits of XRPD data of polycrystalline octreotide at ambient conditions (capillary mode) and selected rH levels (95%, 70%, 60%, 40%, and 30%). The data extend up to ~2.35 Å resolution. They were collected employing a laboratory X-ray powder diffractometer (X’Pert Pro, Malvern Panalytical) equipped with an Anton Paar MHC-trans humidity-temperature chamber [λ = 1.540598 Å, RT]. In each panel, the black and red lines represent the experimental data and the calculated profiles, respectively, while the blue line corresponds to the difference between the experimental and calculated profiles. The vertical bars indicate the Bragg reflections compatible with this space group (<span class="html-italic">P</span>2<sub>1</sub>2<sub>1</sub>2<sub>1</sub>, lattice parameters at ambient conditions: a = 18.608(2) Å, b = 30.254(3) Å, and c = 39.794(6) Å).</p>
Full article ">Figure 2
<p>Upper panel: Pawley fit of the XRPD synchrotron data of octreotide. The data were collected on ID22 at ESRF and extend up to ~ 2.35 Å resolution [λ = 1.3007899(8) Å, RT]. The black, red, and lower green lines represent the experimental data, the calculated pattern, and the difference between the experimental and calculated profiles, respectively. The orange vertical bars correspond to Bragg reflections compatible with this space group (<span class="html-italic">P</span>2<sub>1</sub>2<sub>1</sub>2<sub>1</sub>, a = 18.57744(4) Å, b = 30.17338(6) Å, and c = 39.70590(9) Å). To highlight the enhanced d-spacing resolution, the profile was systematically multiplied by factors of 5 and 16, as indicated in the figure. Lower panel: Magnification of the 2θ range from 4.3° to 5.6°, emphasizing the enhanced angular resolution of the diffraction pattern. The background intensity has been subtracted for clarity.</p>
Full article ">Figure 3
<p>Surface plots of laboratory XRPD data of the octreotide polycrystalline precipitate upon gradual dehydration/rehydration cycles from 95% to 60% rH (<b>left</b>), 95% to 40% rH (<b>middle</b>), and 95% to 30% (<b>right</b>). Alterations of the peak positions and intensities are evident upon gradual dehydration and rehydration cycles. Significant peak shifts become evident upon dehydration, particularly below 70% rH. Upon rehydration and above 75% rH, the sample effectively recovers to its initial state.</p>
Full article ">Figure 4
<p>Evolution of normalized unit-cell parameters upon gradual dehydration and rehydration of the octreotide polycrystalline sample from 95% to 60% rH (<b>upper panel</b>), 40% rH (<b>middle panel</b>), and 30% rH (<b>lower panel</b>). Purple, red, green, and blue symbols correspond to the extracted parameters of the unit-cell volume V, the a axis, the b axis, and the c axis, respectively. The lines are guides to the eye.</p>
Full article ">Figure 5
<p>Magnification of the laboratory XRPD data in the 4–8° 2<span class="html-italic">θ</span> range reveals significant peak shifts at 75% and 70% rH (RT), along with the subsequent recovery of the sample after rehydration.</p>
Full article ">Figure 6
<p>Following a complete dehydration and rehydration cycle, a comparative view of XRPD data at 95% rH reveals a pronounced improvement of the XRPD data in terms of signal-to-noise ratio.</p>
Full article ">Figure 7
<p>Surface plots of XRPD data of octreotide polycrystalline precipitate upon gradual heating/cooling cycles at specific rH levels. A noticeable shift in the diffraction peak positions and intensities is observed at 75% and 65% rH. The latter observation may be attributed more to the effect of humidity than temperature. In addition, the XRPD data collected upon rH variation at ambient temperature described above indicate that at rH lower than 75%, the sample exhibits a slight alteration in terms of unit-cell dimensions, yet crystallinity is maintained. The latter suggests that temperature does not significantly impact the structural integrity of the polycrystalline peptide sample.</p>
Full article ">Figure 8
<p>Evolution of normalized unit-cell parameters upon gradual heating and cooling cycles of the octreotide polycrystalline sample from 294.15 K to 318.15 K at selected rH levels. Turquoise, pink, dark blue, light blue, brown, and green symbols correspond to the rH levels of 95%, 85%, 75%, 65%, 55%, and 45%, respectively. The lines are guides to the eye.</p>
Full article ">Figure 9
<p>Optical microscopy images of the polycrystalline octreotide sample.</p>
Full article ">Figure 10
<p>(<b>a</b>) Empty Kapton foil holder (<b>left</b>) and Kapton foil holder filled with the polycrystalline octreotide sample (<b>right</b>). (<b>b</b>) View of the interior of the humidity chamber containing the multiple-position sample holder. (<b>c</b>) The configuration of the X’Pert Pro diffractometer equipped with the MHC-trans humidity and temperature chamber for in situ XRPD data collection in transmission mode.</p>
Full article ">
18 pages, 46006 KiB  
Article
Exploring the Composition of Egyptian Faience
by Francesca Falcone, Maria Aquilino and Francesco Stoppa
Minerals 2024, 14(6), 586; https://doi.org/10.3390/min14060586 - 31 May 2024
Cited by 1 | Viewed by 801
Abstract
Egyptian Faience, a revolutionary innovation in ancient ceramics, was used for crafting various objects, including amulets, vessels, ornaments, and funerary figurines, like shabtis. Despite extensive research, many aspects of ancient shabti production technology, chemistry and mineralogy remain relatively understudied from the 21st to [...] Read more.
Egyptian Faience, a revolutionary innovation in ancient ceramics, was used for crafting various objects, including amulets, vessels, ornaments, and funerary figurines, like shabtis. Despite extensive research, many aspects of ancient shabti production technology, chemistry and mineralogy remain relatively understudied from the 21st to the 22nd Dynasty, belonging to a recovered 19th-century private collection. The fragments’ origin is tentatively identified in the middle Nile valley in the Luxor area. Our study focused on a modest yet compositionally interesting small collection of shabti fragments to provide information on the glaze’s components and shabti’s core. We found that the core is a quartz and K-feldspars silt blended with an organic component made of plastic resins and vegetable fibres soaked with natron. The studied shabti figurines, after being modelled, dried, and covered with coloured glaze, were subjected to a firing process. Sodium metasilicate and sulphate compounds formed upon contact of the glaze with the silica matrix, forming a shell that holds together the fragile inner matrix. The pigments dissolved in the sodic glaze glass, produced by quartz, K-feldspars, and natron frit, are mainly manganese (Mn) and copper (Cu) compounds. The ratio Cu2O/CaO > 5 produces a blue colour; if <5, the glaze is green. In some cases, Mg and As may have been added to produce a darker brown and an intense blue, respectively. Reaction minerals provided information on the high-temperature firing process that rapidly vitrified the glaze. These data index minerals for the firing temperature of a sodic glaze, reaching up to a maximum of 1050 °C. Full article
Show Figures

Figure 1

Figure 1
<p>An image of a small collection of Egyptian Faience, as it arrived when generously donated to the university by an anonymous donor. The figure displays the fragments deemed representative and selected for archaeometric study, numbered 1–4.</p>
Full article ">Figure 2
<p>Samples were observed under an optical reflected light microscope (4.5X). Images (<b>A</b>), (<b>B</b>), and (<b>C</b>) are related to sample 1; (<b>D</b>), (<b>E</b>), and (<b>F</b>) are related to sample 2; (<b>G</b>), (<b>H</b>), and (<b>I</b>) are related to sample 3; and (<b>L</b>), (<b>M</b>), and (<b>N</b>) are related to sample 4. (<b>A</b>) Sample 1; (<b>B</b>) inset in A showing the coating (glaze) on an altered blue-greenish background and the presence of pores and gaps; and (<b>C</b>) glaze and core in the section of sample 1. (<b>D</b>) Whole sample 2; (<b>E</b>) inset in D showing glaze coating with brown inscriptions; and (<b>F</b>) section of sample 2. (<b>G</b>) Whole sample 3; (<b>H</b>) inset in G; and (<b>I</b>) cross section of samples 3. (<b>L</b>) Sample 4; (<b>M</b>) inset in L showing brown inscriptions; and (<b>N</b>) section of sample 4.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>D</b>) Two examples of shabti [<a href="#B14-minerals-14-00586" class="html-bibr">14</a>] with decorations like those of the analysed fragments. (<b>B</b>) Sample 2, Shabti with a hieroglyphic inscription located in the exact position in the figurine; (<b>C</b>) inscription; (<b>D</b>) mummiform shabti in Faience with a seshed headband tied at the back; and (<b>E</b>) sample 4 with a seshed headband tied at the back.</p>
Full article ">Figure 4
<p>SEM-EDS analyses of core samples were investigated; the red arrow indicates vegetable fibres. (<b>A</b>) The core of sample 1; (<b>B</b>) the core of sample 2; (<b>C</b>) the greyscale imaging of sample 3; and (<b>D</b>) the combined BSE imaging of sample 3. Organic structures allow for the differentiation of cotton fibres from flax. (<b>E</b>) The typical spiral ribbon shape of cotton fibres and (<b>F</b>) the typical “<span class="html-italic">bamboo</span>” shape of flax fibres. Mineral abbreviations in the figure: Cal–calcite; Kfs–K-feldspar; Qz–quartz.</p>
Full article ">Figure 5
<p>XRPD diffraction patterns of the samples (core and glaze). The deep blue diffraction pattern is for sample 1; the red diffraction pattern is for sample 2; the red diffraction pattern is for sample 3; and the green diffraction pattern is for sample 4. Mineral abbreviations are Ata–atacamite, Cal–calcite, Cuv–cuprorivaite, Gp–gypsum, Or–orthoclase, Hem–hematite; Mgt–magnetite; Mnn–manganite; Ttn–titanite; Qz–quartz. The Laurence N. Warr 2021 IMA abbreviation defines Minerals in the figure.</p>
Full article ">Figure 6
<p>Diagrams of the inner glaze composition. (<b>A</b>) Ternary diagram of SiO<sub>2</sub>/Al<sub>2</sub>O<sub>3</sub>+Na<sub>2</sub>O+K<sub>2</sub>O/CaO+MgO+FeO (after Tite, 1997); (<b>B</b>) ternary diagram of SiO<sub>2</sub>/Al<sub>2</sub>O<sub>3</sub>+Na<sub>2</sub>O+K<sub>2</sub>O/CaO+MgO+FeO+Cu<sub>2</sub>O+MnO (plus pigments); (<b>C</b>) binary diagram of Na<sub>2</sub>O vs. SiO<sub>2</sub>, where the regression line represents the degree of natron saturation of the glass in the glaze natron/quartz, with a maximum of 35% natron added; (<b>D</b>) CuO vs. Ca, where the dividing line represent two different compositions referring to high Cu glazes (samples 2–3) and low Cu glazes (samples 1–4). The legend and symbols are in the figure.</p>
Full article ">Figure 7
<p>A spider diagram normalised based on the average glass glaze composition of the three glaze colouration types.</p>
Full article ">Figure 8
<p>SEM images of a polished section: (<b>A</b>) sample 1 and (<b>B</b>) sample 4. The red dotted line highlighted by the red arrows, delineates the reaction shell between the core-matrix and the glaze.</p>
Full article ">
19 pages, 7940 KiB  
Article
Electrospun Nanofibers with Pomegranate Peel Extract as a New Concept for Treating Oral Infections
by Magdalena Paczkowska-Walendowska, Miłosz Ignacyk, Andrzej Miklaszewski, Tomasz Plech, Tomasz M. Karpiński, Jakub Kwiatek, Ewelina Swora-Cwynar, Michał Walendowski and Judyta Cielecka-Piontek
Materials 2024, 17(11), 2558; https://doi.org/10.3390/ma17112558 - 26 May 2024
Cited by 2 | Viewed by 1134
Abstract
Pomegranate peel extract is known for its potent antibacterial, antiviral, antioxidant, anti-inflammatory, wound healing, and probiotic properties, leading to its use in treating oral infections. In the first stage of this work, for the first time, using the Design of Experiment (DoE) approach, [...] Read more.
Pomegranate peel extract is known for its potent antibacterial, antiviral, antioxidant, anti-inflammatory, wound healing, and probiotic properties, leading to its use in treating oral infections. In the first stage of this work, for the first time, using the Design of Experiment (DoE) approach, pomegranate peel extract (70% methanol, temperature 70 °C, and three cycles per 90 min) was optimized and obtained, which showed optimal antioxidant and anti-inflammatory properties. The optimized extract showed antibacterial activity against oral pathogenic bacteria. The second part of this study focused on optimizing an electrospinning process for a combination of polycaprolactone (PCL) and polyvinylpyrrolidone (PVP) nanofibers loaded with the optimized pomegranate peel extract. The characterization of the nanofibers was confirmed by using SEM pictures, XRPD diffractograms, and IR-ATR spectra. The composition of the nanofibers can control the release; in the case of PVP–based nanofibers, immediate release was achieved within 30 min, while in the case of PCL/PVP, controlled release was completed within 24 h. Analysis of the effect of different scaffold compositions of the obtained electrofibers showed that those based on PCL/PVP had better wound healing potential. The proposed strategy to produce electrospun nanofibers with pomegranate peel extract is the first and innovative approach to better use the synergy of biological action of active compounds present in extracts in a patient-friendly pharmaceutical form, beneficial for treating oral infections. Full article
(This article belongs to the Section Advanced Nanomaterials and Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SEM images of nanofibers N1–N6.</p>
Full article ">Figure 1 Cont.
<p>SEM images of nanofibers N1–N6.</p>
Full article ">Figure 2
<p>XRPD diffractograms for extract and nanofibers N1–N6.</p>
Full article ">Figure 3
<p>IR-ATR spectra for PVP, PCL, extract (<b>a</b>), and nanofibers N1–N6 (<b>b</b>).</p>
Full article ">Figure 4
<p>Dissolution profiles of rutin from nanofibers N1, N3, and N5.</p>
Full article ">Figure 5
<p>Mucoadhesive properties of nanofibers N1–N6.</p>
Full article ">Figure 6
<p>Viability of human normal skin fibroblasts (Hs27 cells) exposed to 100 µg/mL of the extract and nanofibers N1–N6 for 24 h. ANOVA statistically analyzed the results of the MTT assay with a post hoc Dunnett’s test. Statistical significance (vs control cells) was designated as *** when <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 7
<p>The wound-healing effect of the extract and nanofibers N1–N6 was observed on two-dimensional cultures of normal human skin fibroblasts (representative images).</p>
Full article ">Figure 8
<p>The wound-healing properties of the extract and nanofibers N1–N6 were examined on Hs27 cells after 24 h incubation. The samples were tested at a concentration of 100 µg/mL. ANOVA statistically analyzed results with a post hoc Tukey’s test. Statistical significance was designated as “*” when <span class="html-italic">p</span> &lt; 0.05, “**” when <span class="html-italic">p</span> &lt; 0.01, “***” when <span class="html-italic">p</span> &lt; 0.001, and “****” when <span class="html-italic">p</span> &lt; 0.0001 (vs. control cells or extract-treated cells); ns—not significant.</p>
Full article ">Figure 9
<p>Principal component analysis (PCA) showing the factor loading plot considering the average diameter of nanofibers (=diameter), production efficiency, percentage of rutin release at 6 h (=dissolution), antioxidant activity (=DPPH), anti-inflammatory activity (=Hyal), mucoadhesive properties, and wound closure after 24 h (=wound closure); N1—nanofibers N1, N3—nanofibers N3, N5—nanofibers N5.</p>
Full article ">
21 pages, 10887 KiB  
Article
Design of Na3MnZr(PO4)3/Carbon Nanofiber Free-Standing Cathodes for Sodium-Ion Batteries with Enhanced Electrochemical Performances through Different Electrospinning Approaches
by Debora Maria Conti, Claudia Urru, Giovanna Bruni, Pietro Galinetto, Benedetta Albini, Chiara Milanese, Silvia Pisani, Vittorio Berbenni and Doretta Capsoni
Molecules 2024, 29(8), 1885; https://doi.org/10.3390/molecules29081885 - 20 Apr 2024
Viewed by 1391
Abstract
The NASICON-structured Na3MnZr(PO4)3 compound is a promising high-voltage cathode material for sodium-ion batteries (SIBs). In this study, an easy and scalable electrospinning approach was used to synthesize self-standing cathodes based on Na3MnZr(PO4)3 loaded [...] Read more.
The NASICON-structured Na3MnZr(PO4)3 compound is a promising high-voltage cathode material for sodium-ion batteries (SIBs). In this study, an easy and scalable electrospinning approach was used to synthesize self-standing cathodes based on Na3MnZr(PO4)3 loaded into carbon nanofibers (CNFs). Different strategies were applied to load the active material. All the employed characterization techniques (X-ray powder diffraction (XRPD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), energy-dispersive X-ray spectroscopy (EDS), thermal gravimetric analysis (TGA), and Raman spectroscopy) confirmed the successful loading. Compared to an appositely prepared tape-cast electrode, Na3MnZr(PO4)3/CNF self-standing cathodes demonstrated an enhanced specific capacity, especially at high C-rates, thanks to the porous conducive carbon nanofiber matrix. Among the strategies applied to load Na3MnZr(PO4)3 into the CNFs, the electrospinning (vertical setting) of the polymeric solution containing pre-synthesized Na3MnZr(PO4)3 powders resulted effective in obtaining the quantitative loading of the active material and a homogeneous distribution through the sheet thickness. Notably, Na3MnZr(PO4)3 aggregates connected to the CNFs, covered their surface, and were also embedded, as demonstrated by TEM and EDS. Compared to the self-standing cathodes prepared with the horizontal setting or dip–drop coating methods, the vertical binder-free electrode exhibited the highest capacity values of 78.2, 55.7, 38.8, 22.2, 16.2, 12.8, 10.3, 9.0, and 8.5 mAh/g at C-rates of 0.05C, 0.1C, 0.2C, 0.5C, 1C, 2C, 5C, 10C, and 20C, respectively, with complete capacity retention at the end of the measurements. It also exhibited a good cycling life, compared to its tape-cast counterpart: it displayed higher capacity retention at 0.2C and 1C, and, after cycling 1000 cycles at 1C, it could be further cycled at 5C, 10C, and 20C. Full article
(This article belongs to the Special Issue Modern Materials in Energy Storage and Conversion)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>A scheme of the samples’ synthesis.</p>
Full article ">Figure 2
<p>XRPD patterns of CNFs, pristine Na<sub>3</sub>MnZr(PO<sub>4</sub>)<sub>3</sub> powder, and self-standing cathodes.</p>
Full article ">Figure 3
<p>Raman spectra of p-MnZr, h-10%MnZr/CNF, h-30%MnZr/CNF, v-30%MnZr/CNF, and dd-MnZr/CNF samples.</p>
Full article ">Figure 4
<p>SEM and TEM images of Na<sub>3</sub>MnZr(PO<sub>4</sub>)<sub>3</sub> powder: (<b>a</b>) 20.42 kX and (<b>b</b>) 100 kX.</p>
Full article ">Figure 5
<p>SEM images of h-10%MnZr/CNF (<b>a</b>,<b>b</b>) surface and (<b>c</b>) cross-section; h-30%MnZr/CNF (<b>d</b>,<b>e</b>) surface and (<b>f</b>) cross-section; v-30%MnZr/CNF (<b>g</b>,<b>h</b>) surface and (<b>i</b>) cross-section; and dd-MnZr/CNF (<b>j</b>,<b>k</b>) surface and (<b>l</b>) cross-section.</p>
Full article ">Figure 6
<p>TEM images at different magnifications (20 kX e 50 kX) of (<b>a</b>–<b>c</b>) h-10%MnZr/CNF, (<b>d</b>–<b>f</b>) h-30%MnZr/CNF, (<b>g</b>–<b>i</b>) v-30%MnZr/CNF, and (<b>j</b>–<b>l</b>) dd-MnZr/CNF samples.</p>
Full article ">Figure 7
<p>SEM image and EDS maps of the different elements for the h-10%MnZr/CNF (<b>a</b>–<b>e</b>) surface and (<b>f</b>–<b>j</b>) its cross-section.</p>
Full article ">Figure 8
<p>SEM image and EDS maps of the different elements for the h-30%MnZr/CNF (<b>a</b>–<b>e</b>) surface and (<b>f</b>–<b>j</b>) its cross-section. The yellow frame indicates the mapped portion.</p>
Full article ">Figure 9
<p>SEM image and EDS maps of the different elements for the v-30%MnZr/CNF (<b>a</b>–<b>e</b>) surface and (<b>f</b>–<b>j</b>) its cross-section. The yellow frame indicates the mapped portion.</p>
Full article ">Figure 10
<p>SEM image and EDS maps of the different elements for the dd-MnZr/CNF (<b>a</b>–<b>e</b>) surface and (<b>f</b>–<b>j</b>) its cross-section. The yellow frame indicates the mapped portion.</p>
Full article ">Figure 11
<p>TGA analysis of p-MnZr (black), h-10%MnZr/CNF (blue), h-30%MnZr/CNF (red) v-30%MnZr/CNF (green), and dd-MnZr/CNF (purple) samples.</p>
Full article ">Figure 12
<p>Charge/discharge cycles at different C-rates of (<b>a</b>) p-MnZr, (<b>b</b>) h-10%MnZr/CNF, (<b>c</b>) h-30%MnZr/CNF, (<b>d</b>) v-30%MnZr/CNF, and (<b>e</b>) dd-MnZr/CNF; (<b>f</b>) comparison of average discharge capacity values for all the samples at different C-rates.</p>
Full article ">Figure 13
<p>Cycling performance of (<b>a</b>) p-MnZr slurry electrode and (<b>b</b>) v-30%MnZr/CNF self-standing cathode.</p>
Full article ">
13 pages, 3032 KiB  
Article
Preparation and Investigation of a Nanosized Piroxicam Containing Orodispersible Lyophilizate
by Petra Party, Sándor Soma Sümegi and Rita Ambrus
Micromachines 2024, 15(4), 532; https://doi.org/10.3390/mi15040532 - 15 Apr 2024
Viewed by 1218
Abstract
Non-steroidal anti-inflammatory piroxicam (PRX) is a poorly water-soluble drug that provides relief in different arthritides. Reducing the particle size of PRX increases its bioavailability. For pediatric, geriatric, and dysphagic patients, oral dispersible systems ease administration. Moreover, fast disintegration followed by drug release and [...] Read more.
Non-steroidal anti-inflammatory piroxicam (PRX) is a poorly water-soluble drug that provides relief in different arthritides. Reducing the particle size of PRX increases its bioavailability. For pediatric, geriatric, and dysphagic patients, oral dispersible systems ease administration. Moreover, fast disintegration followed by drug release and absorption through the oral mucosa can induce rapid systemic effects. We aimed to produce an orodispersible lyophilizate (OL) consisting of nanosized PRX. PRX was solved in ethyl acetate and then sonicated into a poloxamer-188 solution to perform spray-ultrasound-assisted solvent diffusion-based nanoprecipitation. The solid form was formulated via freeze drying in blister sockets. Mannitol and sodium alginate were applied as excipients. Dynamic light scattering (DLS) and nanoparticle tracking analysis (NTA) were used to determine the particle size. The morphology was characterized by scanning electron microscopy (SEM). To establish the crystallinity, X-ray powder diffraction (XRPD) and differential scanning calorimetry (DSC) were used. A disintegration and in vitro dissolution test were performed. DLS and NTA presented a nanosized PRX diameter. The SEM pictures showed a porous structure. PRX became amorphous according to the XRPD and DSC curves. The disintegration time was less than 1 min and the dissolution profile improved. The final product was an innovative anti-inflammatory drug delivery system. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic figure of the preparation method (created using <a href="http://BioRender.com" target="_blank">BioRender.com</a>, accessed on 12 February 2024.).</p>
Full article ">Figure 2
<p>NTA results of the NS (red color shows the distribution of the particles, blue color shows the size of the particles at exact points.</p>
Full article ">Figure 3
<p>SEM images of the NS (<b>a</b>) and the OL (<b>b</b>).</p>
Full article ">Figure 4
<p>DSC results of the initial materials, the NS, and the OL.</p>
Full article ">Figure 5
<p>XRPD results of the initial materials, the NS, and the OL.</p>
Full article ">Figure 6
<p>FTIR results of two initial materials: the PM and the OL.</p>
Full article ">Figure 7
<p>Dissolution of the initial drug and the OL. Data are means ± SD (<span class="html-italic">n</span> = 3 independent measurements).</p>
Full article ">
Back to TopTop