[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

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,915)

Search Parameters:
Keywords = core–shell

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 5253 KiB  
Article
Targeted PHA Microsphere-Loaded Triple-Drug System with Sustained Drug Release for Synergistic Chemotherapy and Gene Therapy
by Shuo Wang, Chao Zhang, Huandi Liu, Xueyu Fan, Shuangqing Fu, Wei Li and Honglei Zhang
Nanomaterials 2024, 14(20), 1657; https://doi.org/10.3390/nano14201657 (registering DOI) - 16 Oct 2024
Viewed by 282
Abstract
The combination of paclitaxel (PTX) with other chemotherapy drugs (e.g., gemcitabine, GEM) or genetic drugs (e.g., siRNA) has been shown to enhance therapeutic efficacy against tumors, reduce individual drug dosages, and prevent drug resistance associated with single-drug treatments. However, the varying solubility of [...] Read more.
The combination of paclitaxel (PTX) with other chemotherapy drugs (e.g., gemcitabine, GEM) or genetic drugs (e.g., siRNA) has been shown to enhance therapeutic efficacy against tumors, reduce individual drug dosages, and prevent drug resistance associated with single-drug treatments. However, the varying solubility of chemotherapy drugs and genetic drugs presents a challenge in co-delivering these agents. In this study, nanoparticles loaded with PTX were prepared using the biodegradable polymer material poly(3-hydroxybutyrate-co-3-hydroxyhexanoate) (PHBHHx). These nanoparticles were surface-modified with target proteins (Affibody molecules) and RALA cationic peptides to create core-shell structured microspheres with targeted and cationic functionalization. A three-drug co-delivery system (PTX@PHBHHx-ARP/siRNAGEM) were developed by electrostatically adsorbing siRNA chains containing GEM onto the microsphere surface. The encapsulation efficiency of PTX in the nanodrug was found to be 81.02%, with a drug loading of 5.09%. The chemogene adsorption capacity of siRNAGEM was determined to be 97.3%. Morphological and size characterization of the nanodrug revealed that PTX@PHBHHx-ARP/siRNAGEM is a rough-surfaced microsphere with a particle size of approximately 150 nm. This nanodrug exhibited targeting capabilities toward BT474 cells with HER2 overexpression while showing limited targeting ability toward MCF-7 cells with low HER2 expression. Results from the MTT assay demonstrated that PTX@PHBHHx-ARP/siRNAGEM exhibits high cytotoxicity and excellent combination therapy efficacy compared to physically mixed PTX/GEM/siRNA. Additionally, Western blot analysis confirmed that siRNA-mediated reduction of Bcl-2 expression significantly enhanced cell apoptosis mediated by PTX or GEM in tumor cells, thereby increasing cell sensitivity to PTX and GEM. This study presents a novel targeted nanosystem for the co-delivery of chemotherapy drugs and genetic drugs. Full article
Show Figures

Figure 1

Figure 1
<p>SEM (<b>a</b>), TEM (<b>b</b>), and FTIR (<b>c</b>) analysis.</p>
Full article ">Figure 2
<p>Release curves of PTX and siRNA<sub>GEM</sub> in PTX@PHBHHx-ARP/siRNA<sub>GEM</sub>. (<b>a</b>) The PTX drug release curve. (<b>b</b>) The release curve of siRNA<sub>GEM</sub>.</p>
Full article ">Figure 3
<p>Drug uptake by BT474 and MCF-7 cells. (<b>a</b>) CLSM analysis of BT474 cells treated with PTX@PHBHHx-ARP/siRNA<sub>GEM</sub> for different times. Panel (<b>b</b>) CLSM analysis of MCF-7 cells treated with PTX@PHBHHx-ARP/siRNA<sub>GEM</sub> for different times. Scale bar: 20 μm.</p>
Full article ">Figure 4
<p>Survival of BT474 and MCF-7 cells treated with PTX, GEM, physical mixture of PTX/GEM/siRNA and PTX@PHBHHx-ARP/siRNA<sub>GEM</sub>. (<b>a</b>) The survival rates of the two cell types after 48 h incubation with PTX monotherapy. (<b>b</b>) The survival rates of the two cell types after 48 h incubation with GEM monotherapy. (<b>c</b>) The survival rates of the two cell types after 48 h of incubation with the physical mixture of PTX/GEM/siRNA. (<b>d</b>) The survival rates of the two cell types after 48 h incubation with PTX@PHBHHx-ARP/siRNA<sub>GEM</sub>.</p>
Full article ">Figure 5
<p>Western blot of Bcl-2 in (<b>a</b>) BT474 cells and (<b>b</b>) MCF-7 cells treated with free PTX, GEM, PTX/GEM/siRNA, and PTX@PHBHHx-ARP/siRNA<sub>GEM</sub>, respectively. (<b>c</b>) Significance analysis of Bcl-2 in BT474 cells. (<b>d</b>) Significance analysis of Bcl-2 in MCF-7 cells. Statistical analysis: * <span class="html-italic">p</span> &lt; 0.05, **** <span class="html-italic">p</span> &lt; 0.0001 and ns (No significant difference).</p>
Full article ">Figure 6
<p>Scatter plots of apoptosis in (<b>a</b>) MCF-7 cells and (<b>b</b>) BT474 cells treated with free PTX, GEM, PTX/GEM/siRNA, and PTX@PHBHHx-ARP/siRNA<sub>GEM</sub>, respectively.</p>
Full article ">Scheme 1
<p>The assembly process of PTX@PHBHHx-ARP/siRNA<sub>GEM</sub> and their synergistic cancer therapy.</p>
Full article ">
12 pages, 4907 KiB  
Article
Multi-Wavelength Excitable Multicolor Upconversion and Ratiometric Luminescence Thermometry of Yb3+/Er3+ Co-Doped NaYGeO4 Microcrystals
by Hui Zeng, Yangbo Wang, Xiaoyi Zhang, Xiangbing Bu, Zongyi Liu and Huaiyong Li
Molecules 2024, 29(20), 4887; https://doi.org/10.3390/molecules29204887 (registering DOI) - 15 Oct 2024
Viewed by 197
Abstract
Excitation wavelength controllable lanthanide upconversion allows for real-time manipulation of luminescent color in a composition-fixed material, which has been proven to be conducive to a variety of applications, such as optical anti-counterfeiting and information security. However, current available materials highly rely on the [...] Read more.
Excitation wavelength controllable lanthanide upconversion allows for real-time manipulation of luminescent color in a composition-fixed material, which has been proven to be conducive to a variety of applications, such as optical anti-counterfeiting and information security. However, current available materials highly rely on the elaborate core–shell structure in order to ensure efficient excitation-dependent energy transfer routes. Herein, multicolor upconversion luminescence in response to both near-infrared I and near-infrared II (NIR-I and NIR-II) excitations is realized in a novel but simple NaYGeO4:Yb3+/Er3+ phosphor. The remarkably enhanced red emission ratio under 1532 nm excitation, compared with that under 980 nm excitation, could be attributed to the Yb3+-mediated cross-relaxation energy transfers. Moreover, multi-wavelength excitable temperature-dependent (295–823 K) upconversion luminescence realizes a ratiometric thermometry relying on the thermally coupled levels (TCLs) of Er3+. Detailed investigations demonstrate that changing excitation wavelength makes little difference for the performances of TCL-based ratiometric thermometry of NaYGeO4:Yb3+/Er3+. These findings gain more insights to manipulate cross-relaxations for excitation controllable upconversion in single activator doped materials and benefit the cognition of the effect of excitation wavelength on ratiometric luminescence thermometry. Full article
(This article belongs to the Special Issue Rare Earth Based Luminescent Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Crystal structure of NaYGeO<sub>4</sub> and the coordination polyhedra of NaO<sub>6</sub>, YO<sub>6</sub>, and GeO<sub>4</sub>. (<b>b</b>) XRD patterns of NaYGeO<sub>4</sub>:xYb<sup>3+</sup>/2%Er<sup>3+</sup> microcrystals, x = 2–48%. The bar-like diffraction patterns at the bottom represent the standard data of orthorhombic NaYGeO<sub>4</sub> (PDF#88–1177). (<b>c</b>) Unit cell volume calculated from Rietveld refinement results as a function of Yb<sup>3+</sup> doping concentration. (<b>d</b>) XPS spectrum, (<b>e</b>) SEM image, and (<b>f</b>) EDS spectrum of NaYGeO<sub>4</sub>:18%Yb<sup>3+</sup>/2%Er<sup>3+</sup>. The inset in (<b>d</b>) is the high-resolution XPS spectrum in the range of 165–190 eV. (<b>g</b>) Elemental mappings corresponding to the SEM image in (<b>e</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) Upconversion emission spectra of NaYGeO<sub>4</sub>:18%Yb<sup>3+</sup>/2%Er<sup>3+</sup>, under 980 and 1532 nm laser excitation. The integral intensity evolutions of 515–570, 634–705, and 783–826 nm emissions for NaYGeO<sub>4</sub>: xYb<sup>3+</sup>/2%Er<sup>3+</sup> at increased Yb<sup>3+</sup> concentrations, under (<b>b</b>) 980 and (<b>c</b>) 1532 nm excitation. (<b>d</b>) The upconversion red/green ratios and (<b>e</b>) luminescence photographs of NaYGeO<sub>4</sub>:xYb<sup>3+</sup>/2%Er<sup>3+</sup> at increased Yb<sup>3+</sup> concentrations under 980 and 1532 nm excitation.</p>
Full article ">Figure 3
<p>Upconversion emission spectra of NaYGeO<sub>4</sub>:18%Yb<sup>3+</sup>/2%Er<sup>3+</sup> and NaYGeO<sub>4</sub>:2%Er<sup>3+</sup>, under (<b>a</b>) 980 and (<b>b</b>) 1532 nm excitation. Schematic upconversion luminescence mechanisms of NaYGeO<sub>4</sub>:Yb<sup>3+</sup>/Er<sup>3+</sup> with (<b>c</b>) 980 and (<b>d</b>) 1532 nm excitation. Decay curves of NaYGeO<sub>4</sub>: xYb<sup>3+</sup>/2%Er<sup>3+</sup> under 1532 nm excitation at (<b>e</b>) 558 and (<b>f</b>) 660 nm emissions; the insets show the calculated lifetimes as a function of Yb<sup>3+</sup> concentration.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>d</b>) Upconversion luminescence spectra, (<b>b</b>,<b>e</b>) normalized green upconversion spectra, and (<b>c</b>,<b>f</b>) calculated Ln LIR (<span class="html-italic">I</span><sub>532</sub>/<span class="html-italic">I</span><sub>558</sub>) of NaYGeO<sub>4</sub>:18%Yb<sup>3+</sup>/2%Er<sup>3+</sup> in the temperature range of 295–823 K, under (<b>a</b>–<b>c</b>) 980 and (<b>d</b>–<b>f</b>) 1532 nm excitation.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>d</b>) Absolute sensitivity, <span class="html-italic">S<sub>a</sub></span>, and relative sensitivity, <span class="html-italic">S<sub>r</sub></span>. (<b>b</b>,<b>e</b>) Temperature uncertainty <span class="html-italic">δT</span> relying on LIR (<span class="html-italic">I</span><sub>532</sub>/<span class="html-italic">I</span><sub>558</sub>) of NaYGeO<sub>4</sub>:18%Yb<sup>3+</sup>/2%Er<sup>3+</sup> at different temperatures. (<b>c</b>,<b>f</b>) LIR (<span class="html-italic">I</span><sub>532</sub>/<span class="html-italic">I</span><sub>558</sub>) at selected temperatures for two heating–cooling cycles between 323 and 823 K. Under (<b>a</b>–<b>c</b>) 980 and (<b>d</b>–<b>f</b>) 1532 nm excitation.</p>
Full article ">
13 pages, 3069 KiB  
Article
Sub-10 nm PdNi@PtNi Core–Shell Nanoalloys for Efficient Ethanol Electro-Oxidation
by Qian Su and Lei Yu
Molecules 2024, 29(20), 4853; https://doi.org/10.3390/molecules29204853 - 13 Oct 2024
Viewed by 432
Abstract
By controlling the structure and composition of Pt-based nanoalloys, the ethanol oxidation reaction (EOR) performances of Pt alloy catalysts can be effectively improved. Herein, we successfully synthesis sub-10 nm PdNi@PtNi nanoparticles (PdNi@PtNi NPs) with a core–shell structure by a one-pot method. The sub [...] Read more.
By controlling the structure and composition of Pt-based nanoalloys, the ethanol oxidation reaction (EOR) performances of Pt alloy catalysts can be effectively improved. Herein, we successfully synthesis sub-10 nm PdNi@PtNi nanoparticles (PdNi@PtNi NPs) with a core–shell structure by a one-pot method. The sub 10 nm core–shell nanoparticles possess more effective atoms and exhibit a synergistic effect which can lead to a shift in the d-band center and alter binding energies toward adsorbates. Due to the synergistic effect and unique core–shell structure, the PdNi@PtNi NP catalysts exhibit excellent electrocatalytic performance for ethanol oxidation reactions in alkaline, achieving 9.30 times more mass activity and 7.05 times more specific activity that of the state-of-the-art Pt/C catalysts. Moreover, the stability of PdNi@PtNi NPs was also greatly improved over PtNi nanoparticles, PtPd nanoparticles, and commercial Pt/C. This strategy provides a new idea for improving the electrocatalytic performance of Pt-based catalysts for EORs. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Illustrations of the synthesis of PdNi@PtNi NPs. (<b>b</b>–<b>d</b>) HR-TEM images of PdNi@PtNi NPs. Inset in b shows the diameter distribution of individual nanoparticles in PdNi@PtNi NPs. (<b>e</b>–<b>i</b>) HAADF-STEM-EDS mapping images of PdNi@PtNi NPs.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns of PdNi@PtNi NPs, PtNi NPs, and PtPd NPs. (<b>b</b>) Ni 2p XPS spectra of the catalysts. (<b>c</b>) Pt 3d XPS spectra of the catalysts. (<b>d</b>) Pd 4f XPS spectra of the catalysts.</p>
Full article ">Figure 3
<p>(<b>a</b>) CV curves for PdNi@PtNi NP, PtNi NP, and PtPd NP catalysts in a 1 M KOH solution at a scan rate of 50 mV⋅s<sup>-1</sup>. (<b>b</b>) Specific current densities of the catalysts; (<b>c</b>) onset potential of the catalysts; (<b>d</b>) specific activities and mass activities of the catalysts; (<b>e</b>) i-t curves of the catalysts; and (<b>f</b>) activity decay on different catalysts during the CV cycles.</p>
Full article ">Figure 4
<p>CO stripping test of (<b>a</b>) PdNi@PtNi NPs, (<b>b</b>) PtPd NPs, (<b>c</b>) PtNi NPs, and (<b>d</b>) Pt/C.</p>
Full article ">Figure 5
<p>Reaction pathways for EORs at the Pt active sites.</p>
Full article ">
13 pages, 3025 KiB  
Review
Active Sites on the CuCo Catalyst in Higher Alcohol Synthesis from Syngas: A Review
by Chun Han, Jing Liu, Le Li, Zeyu Peng, Luyao Wu, Jiarong Hao and Wei Huang
Molecules 2024, 29(20), 4855; https://doi.org/10.3390/molecules29204855 - 13 Oct 2024
Viewed by 335
Abstract
Higher alcohol synthesis through the Fischer–Tropsch (F–T) process was considered a promising route for the efficient utilization of fossil resources could be achieved. The CuCo catalysts were proven to be efficient candidates and attracted much interest. Great efforts have been made to investigate [...] Read more.
Higher alcohol synthesis through the Fischer–Tropsch (F–T) process was considered a promising route for the efficient utilization of fossil resources could be achieved. The CuCo catalysts were proven to be efficient candidates and attracted much interest. Great efforts have been made to investigate the active sites and mechanisms of CuCo catalysts. However, the industrialized application of CuCo catalysts in this process was still hindered. The poor stability of this catalyst was one of the main reasons. This short review summarized the recent development of active sites on the CuCo catalysts for higher alcohol synthesis, including CuCo alloy particles, CuCo core–shell particles, and unsaturated particles. The complex active sites and their continual changes during the reaction led to the poor stability of the catalysts. The effect of active sites on catalytic performance was discussed. Furthermore, the key factors in fabricating stable CuCo catalysts were proposed. Finally, reasonable proposals were proposed for designing efficient and stable CuCo catalysts in higher alcohol synthesis. Full article
(This article belongs to the Section Applied Chemistry)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of (<b>a</b>) the calcinated and (<b>b</b>) reduced catalysts [<a href="#B38-molecules-29-04855" class="html-bibr">38</a>]. Copyright (2018) American Chemical Society.</p>
Full article ">Figure 2
<p>Schematic diagram of the mechanism of ethanol formation via the synergistic effect of Cu<sup>+</sup>-Co<sup>0</sup> [<a href="#B49-molecules-29-04855" class="html-bibr">49</a>]. H* represented the dissociated H, CO* represented the non-dissociative adsorbed CO. Copyright (2019) Elsevier.</p>
Full article ">Figure 3
<p>Schematic diagram of the mechanism of HA synthesis via the synergistic effect of CoO<sub>x</sub>-Co<sup>0</sup> [<a href="#B62-molecules-29-04855" class="html-bibr">62</a>]. H* represented the dissociated H species. CO* represented the non-dissociative adsorbed CO species. CH<sub>x</sub>* represented the intermediates formed by the hydrogenation of dissociative adsorbed CO. Copyright (2018) American Chemical Society.</p>
Full article ">Figure 4
<p>Structure evolution of CuCo alloy in HA synthesis. (<b>a</b>) The stability test of CuCo catalysts during 800 h [<a href="#B86-molecules-29-04855" class="html-bibr">86</a>]; (<b>b</b>) conversion of CO and selectivity of CO<sub>2</sub>, hydrocarbons, and total alcohols. (<b>c</b>) Distribution of alcohols. Copyright (2017) American Chemical Society.</p>
Full article ">Figure 5
<p>Catalytic performance of CuCo catalysts varying Cu/Co ratio. (<b>a</b>) Catalytic performance of CuCo catalysts varying Cu/Co ratio, (<b>b</b>) Schematic diagram of “Seesaw” phenomenon. (Note: the red and black arrow in (<b>b</b>) represents the decrease and increase in Cu/Co ratio, respectively).</p>
Full article ">Figure 6
<p>Effect of Cu/Co ratio on the active sites over Cu-Co catalysts [<a href="#B17-molecules-29-04855" class="html-bibr">17</a>]. Copyright (2014) John Wiley and Sons.</p>
Full article ">
11 pages, 3561 KiB  
Article
Fabrication and Characterization of CeO2-Doped Yttria-Stabilized ZrO2 Composite Particles
by Young Seo Kim, Yoon-Suk Oh and Gye Seok An
Processes 2024, 12(10), 2202; https://doi.org/10.3390/pr12102202 - 10 Oct 2024
Viewed by 374
Abstract
The present study focuses on the fabrication and characterization of cerium oxide (CeO2)-doped yttria-stabilized zirconia (YSZ) composite particles, aiming to enhance the durability of thermal barrier coatings (TBCs) in high-temperature applications such as gas turbines and aircraft engines. The incorporation of [...] Read more.
The present study focuses on the fabrication and characterization of cerium oxide (CeO2)-doped yttria-stabilized zirconia (YSZ) composite particles, aiming to enhance the durability of thermal barrier coatings (TBCs) in high-temperature applications such as gas turbines and aircraft engines. The incorporation of CeO2 into the YSZ matrix was motivated by the need to address the limitations of YSZ coatings, particularly their phase transformation and thermal degradation at temperatures exceeding 1300 °C. The synthesis of a composite with a core–shell structure, where CeO2 is doped into YSZ particles, was pursued to improve the thermal stability and reduce the thermal conductivity of the material. The fabrication process involved surface treatment of YSZ particles with HCl and NH4OH to enhance their dispersion characteristics, followed by the adsorption of CeO2 nanoparticles precipitated from Ce precursors. The study revealed a reduction in the average particle size and improved the dispersion stability of the surface-treated YSZ. Notably, base-treated YSZ exhibited increased CeO2 adsorption due to the strong interaction between Ce ions and surface hydroxyl groups. The successful formation of the YSZ@CeO2 core–shell structure was confirmed through XRD, HR-TEM, and SAED analyses. The study suggest that base-treated YSZ@CeO2 composites have the potential to extend the operating life and improve the performance of TBCs under extreme temperature conditions, which may contribute to the development of more resilient thermal barrier systems. Full article
(This article belongs to the Section Materials Processes)
Show Figures

Figure 1

Figure 1
<p>TEM images of (<b>a</b>) raw YSZ, (<b>b</b>) acid-treated YSZ, and (<b>c</b>) base-treated YSZ. SEM images of (<b>d</b>) raw YSZ, (<b>e</b>) acid-treated YSZ, and (<b>f</b>) base-treated YSZ.</p>
Full article ">Figure 2
<p>XRD spectra of raw YSZ and YSZ surface treated with acid and base solutions.</p>
Full article ">Figure 3
<p>XPS spectra of raw YSZ and YSZ surface treated with acid and base.</p>
Full article ">Figure 4
<p>FT−IR spectra of the raw YSZ and YSZ surface treated with acid and base solutions.</p>
Full article ">Figure 5
<p>Particle-size distribution of raw YSZ and YSZ surface treated using acid and base solutions.</p>
Full article ">Figure 6
<p>TEM images of YSZ before and after surface treatment, showing the crystallography of YSZ@CeO<sub>2</sub>: (<b>a</b>) raw YSZ, (<b>b</b>) acid-treated YSZ@CeO<sub>2</sub>, and (<b>c</b>) base-treated YSZ@CeO<sub>2</sub>. (<b>d</b>) XRD spectra of YSZ and YSZ@CeO<sub>2</sub>. (<b>e</b>) SAED images of dot patterns, representing YSZ, and ring patterns, representing CeO<sub>2</sub>. (<b>f</b>) HR-TEM images of YSZ@CeO<sub>2</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) STEM image of acid-treated YSZ@CeO<sub>2</sub> nanoparticles of EDS elemental mapping with separate maps shown for O, Y, Zr, and Ce and (<b>b</b>) base-treated YSZ@CeO<sub>2</sub> nanoparticles of EDS elemental mapping with separate maps shown for O, Y, Zr, and Ce.</p>
Full article ">
40 pages, 5303 KiB  
Review
Advances in Electrospun Poly(ε-caprolactone)-Based Nanofibrous Scaffolds for Tissue Engineering
by Karla N. Robles, Fatima tuz Zahra, Richard Mu and Todd Giorgio
Polymers 2024, 16(20), 2853; https://doi.org/10.3390/polym16202853 - 10 Oct 2024
Viewed by 623
Abstract
Tissue engineering has great potential for the restoration of damaged tissue due to injury or disease. During tissue development, scaffolds provide structural support for cell growth. To grow healthy tissue, the principal components of such scaffolds must be biocompatible and nontoxic. Poly(ε-caprolactone) (PCL) [...] Read more.
Tissue engineering has great potential for the restoration of damaged tissue due to injury or disease. During tissue development, scaffolds provide structural support for cell growth. To grow healthy tissue, the principal components of such scaffolds must be biocompatible and nontoxic. Poly(ε-caprolactone) (PCL) is a biopolymer that has been used as a key component of composite scaffolds for tissue engineering applications due to its mechanical strength and biodegradability. However, PCL alone can have low cell adherence and wettability. Blends of biomaterials can be incorporated to achieve synergistic scaffold properties for tissue engineering. Electrospun PCL-based scaffolds consist of single or blended-composition nanofibers and nanofibers with multi-layered internal architectures (i.e., core-shell nanofibers or multi-layered nanofibers). Nanofiber diameter, composition, and mechanical properties, biocompatibility, and drug-loading capacity are among the tunable properties of electrospun PCL-based scaffolds. Scaffold properties including wettability, mechanical strength, and biocompatibility have been further enhanced with scaffold layering, surface modification, and coating techniques. In this article, we review nanofibrous electrospun PCL-based scaffold fabrication and the applications of PCL-based scaffolds in tissue engineering as reported in the recent literature. Full article
(This article belongs to the Special Issue Functional Polymers for Drug Delivery System II)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of poly(ε-caprolactone).</p>
Full article ">Figure 2
<p>Publications from the last two decades listed on Web of Science containing keywords “electrospinning” and “tissue engineering” and “PCL”, “electrospinning”, and “tissue engineering”.</p>
Full article ">Figure 3
<p>Schematic of fabrication of a nanofibrous scaffold with the electrospinning technique. Electrospinning set-up with syringe pump, steel emitter, voltage supply, and ground collector.</p>
Full article ">Figure 4
<p>Types of PCL-based nanofibers fabricated with electrospinning.</p>
Full article ">Figure 5
<p>Types of emergent scaffolds fabricated from electrospun pristine PCL nanofibers, blend or composite nanofibers, and dual or multi-layer nanofibers.</p>
Full article ">Figure 6
<p>Overview of biomaterials commonly blend-electrospun with PCL to fabricate composite PCL-based scaffolds.</p>
Full article ">Figure 7
<p>Overview of biomaterials commonly blend-electrospun with PCL and their relative influence on scaffold bioactivity and mechanical strength.</p>
Full article ">Figure 8
<p>Overview of scaffold post-electrospinning processing techniques for enhanced functionality or biocompatibility.</p>
Full article ">Figure 9
<p>In vivo wound healing of gelatine-coated PCL nanofibers as well as gelatine-coated drug-loaded PCL nanofibers [<a href="#B120-polymers-16-02853" class="html-bibr">120</a>].</p>
Full article ">Figure 10
<p>Burn wound healing progress at days 3, 7, and 14 for PCL, PCL/GLT, and PCL/GLT/LSLE electrospun fibrous mats [<a href="#B121-polymers-16-02853" class="html-bibr">121</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Schematic of the fabrication process of the PCL/TiO<sub>2</sub>@Cotton Janus membrane. (<b>b</b>) SEM image of the cross-section of the Janus membrane (<b>top</b>). Photographs of water droplets (~30 μL) on the TiO<sub>2</sub>@Cotton layer (<b>middle</b>) and the reversed PCL fibrous layer (<b>bottom</b>). Water droplets were colored by dyes (Acid Violet 7, Sunset Yellow FCF, Alkali Blue 70). (<b>c</b>) SEM images of pristine cotton fabric and its partially enlarged cotton yarn; (<b>d</b>) single cotton yarn and its partial enlarged image; (<b>e</b>) TiO<sub>2</sub>@Cotton yarn and its partial enlarged image. (<b>f</b>–<b>i</b>) SEM images, partial enlarged SEM images, and statistically average fiber diameters corresponding to PCL-7.5, PCL-10, PCL-12.5, and PCL-15, respectively [<a href="#B126-polymers-16-02853" class="html-bibr">126</a>].</p>
Full article ">Figure 12
<p>Cumulative drug release profile of different PIP-loaded PCL formulations [<a href="#B129-polymers-16-02853" class="html-bibr">129</a>].</p>
Full article ">Figure 13
<p>SEM micrographs of different layers of multilayered electrospun fibrous structures. The surface morphology and section view of MGPCL (<b>a</b>,<b>b</b>), The microstructure of SKMGPCL (<b>c</b>,<b>d</b>) and MSKPCL (<b>e</b>,<b>f</b>) [<a href="#B140-polymers-16-02853" class="html-bibr">140</a>].</p>
Full article ">
17 pages, 4023 KiB  
Article
New Technology of Rumen-Protected Bypass Lysine Encapsulated in Lipid Matrix of Beeswax and Carnauba Wax and Natural Tannin Blended for Ruminant Diets
by Claudiney Felipe Almeida Inô, José Morais Pereira Filho, Roberto Matheus Tavares de Oliveira, Juliana Felipe Paula de Oliveira, Edson Cavalcanti da Silva Filho, Ariane Maria da Silva Santos Nascimento, Ronaldo Lopes Oliveira, Romilda Rodrigues do Nascimento, Kevily Henrique de Oliveira Soares de Lucena and Leilson Rocha Bezerra
Animals 2024, 14(19), 2895; https://doi.org/10.3390/ani14192895 - 8 Oct 2024
Viewed by 508
Abstract
Tannins are compounds present in forage plants that, in small quantities in the diet of ruminants, produce protein complexes that promote passage through the rumen and use in the intestine. This study tested the hypothesis that beeswax (BW) and carnauba wax (CW) lipid [...] Read more.
Tannins are compounds present in forage plants that, in small quantities in the diet of ruminants, produce protein complexes that promote passage through the rumen and use in the intestine. This study tested the hypothesis that beeswax (BW) and carnauba wax (CW) lipid matrices are effective encapsulants for creating bypass lysine (Lys) for ruminants, with tannin extracted from the Mimosa tenuiflora hay source enhancing material protection. Microencapsulated systems were made using the fusion–emulsification technique with a 2:1 shell-to-core ratio and four tannin levels (0%, 1%, 2%; 3%). The following eight treatments were tested: BWLys0%, BWLys1%, BWLys2%, BWLys3%, CWLys0%, CWLys1%, CWLys2%, and CWLys3%. Tannin inclusion improved microencapsulation yield and efficiency. CWLys3% had the highest microencapsulation efficiency and retained Lys. Lysine in BW and CW matrices showed higher thermal stability than in its free form. Material retention was greater in BW than CW. Rumen pH and temperature remained unaffected, indicating that BW and CW as the shell and tannin as the adjuvant are efficient encapsulants for Lys bypass production. The formulation CWLys3% is recommended as it is more efficient in protecting the lysin amino acid from rumen degradation. Full article
Show Figures

Figure 1

Figure 1
<p>Efficiency and yield of lysine bypass encapsulated with beeswax (BW) and carnauba wax (CW) with tannin levels of 0, 1, 2 and 3% produced by the fusion–emulsification method.</p>
Full article ">Figure 2
<p>Differential scanning calorimetry (DSC) curve of lysine, beeswax (<b>a</b>) and carnauba wax (<b>b</b>) and lysin bypass (BWLys and CWLys) with natural tannin levels of 0, 1, 2 and 3% produced by the fusion–emulsification method.</p>
Full article ">Figure 3
<p>(<b>A</b>) Water activity (WA), (<b>B</b>) dry matter (DM), and (<b>C</b>) crude protein (CP) of lysin bypass encapsulated into beeswax and carnauba wax (BWLys and CWLys) with natural tannin levels of 0, 1, 2 and 3% produced by the fusion–emulsification method. Different letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05; same letters indicated no significant differences (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 4
<p>Scanning electron micrographs: (<b>A</b>) carnauba wax; (<b>B</b>) beeswax; (<b>C</b>) lysine; (<b>D</b>) <span class="html-italic">Mimosa tenuiflora</span> tannic extract; (<b>E</b>) carnauba wax + lysine; (<b>F</b>) beeswax + lysine; (<b>G</b>) beeswax + lysine + <span class="html-italic">Mimosa tenuiflora</span> tannic extract; (<b>H</b>) carnauba wax + lysine + <span class="html-italic">Mimosa tenuiflora</span> tannic extract.</p>
Full article ">Figure 5
<p>Degradation and retention of dry matter (DM) of encapsulates as a function of the level of tannin, regardless of the type of wax and incubation time in the DaisyII ANKOM. Different letters indicate significant differences at <span class="html-italic">p</span> &lt; 0.05; same letters indicated no significant differences (<span class="html-italic">p</span> &gt; 0.05).</p>
Full article ">Figure 6
<p>Degradation and retention of dry matter (DM) of lysin bypass encapsulated products as a function of tannin levels (<b>a</b>), and in situ incubation time (<b>b</b>).</p>
Full article ">Figure 7
<p>Average nitrogen (N) and crude protein (CP) of lysin bypass encapsulated products as a function of tannin level (<b>a</b>) and in situ incubation time (<b>b</b>).</p>
Full article ">Figure 8
<p>Average rumen (<b>a</b>) pH and (<b>b</b>) temperature (°C) and at different incubation times in the fistulated animal, independent on the type of wax (beeswax and carnauba) as a function of the level (0; 1; 2 and 3%) of tannin added in material as an adjuvant.</p>
Full article ">
19 pages, 3026 KiB  
Article
Stable Polymer-Lipid Hybrid Nanoparticles Based on mcl-Polyhydroxyalkanoate and Cationic Liposomes for mRNA Delivery
by Sergey M. Shishlyannikov, Ilya N. Zubkov, Vera V. Vysochinskaya, Nina V. Gavrilova, Olga A. Dobrovolskaya, Ekaterina A. Elpaeva, Mikhail A. Maslov and Andrey Vasin
Pharmaceutics 2024, 16(10), 1305; https://doi.org/10.3390/pharmaceutics16101305 - 7 Oct 2024
Viewed by 688
Abstract
Background/Objectives: The development of polymer–lipid hybrid nanoparticles (PLNs) is a promising area of research, as it can help increase the stability of cationic lipid carriers. Hybrid PLNs are core–shell nanoparticle structures that combine the advantages of both polymer nanoparticles and liposomes, especially in [...] Read more.
Background/Objectives: The development of polymer–lipid hybrid nanoparticles (PLNs) is a promising area of research, as it can help increase the stability of cationic lipid carriers. Hybrid PLNs are core–shell nanoparticle structures that combine the advantages of both polymer nanoparticles and liposomes, especially in terms of their physical stability and biocompatibility. Natural polymers such as polyhydroxyalkanoate (PHA) can be used as a matrix for the PLNs’ preparation. Methods: In this study, we first obtained stable cationic hybrid PLNs using a cationic liposome (CL) composed of a polycationic lipid 2X3 (1,26-bis(cholest-5-en-3β-yloxycarbonylamino)-7,11,16,20-tetraazahexacosane tetrahydrochloride), helper lipid DOPE (1,2-dioleoyl-sn-glycero-3-phosphoethanolamine), and the hydrophobic polymer mcl-PHA, which was produced by the soil bacterium Pseudomonas helmantisensis P1. Results: The new polymer-lipid carriers effectively encapsulated and delivered model mRNA-eGFP (enhanced green fluorescent protein mRNA) to BHK-21 cells. We then evaluated the role of mcl-PHA in increasing the stability of cationic PLNs in ionic solutions using dynamic light scattering data, electrophoretic mobility, and transmission electron microscopy techniques. Conclusions: The results showed that increasing the concentration of PBS (phosphate buffered saline) led to a decrease in the stability of the CLs. At high concentrations of PBS, the CLs aggregate. In contrast, the presence of isotonic PBS did not result in the aggregation of PLNs, and the particles remained stable for 120 h when stored at +4 °C. The obtained results show that PLNs hold promise for further in vivo studies on nucleic acid delivery. Full article
(This article belongs to the Special Issue Polymer-Based Delivery System)
Show Figures

Figure 1

Figure 1
<p>Reactor for the synthesis of PLNs. Reaction tube and the sonicator probe are placed in a beaker with water. During the synthesis and under sonication, the organic solvent (<span class="html-italic">n</span>-hexane) is removed by evaporation in argon stream.</p>
Full article ">Figure 2
<p>Analysis of mRNA binding with PLNs by capillary electrophoresis. Control RNA—control mRNA samples without PLNs containing 50 ng of mRNA. PHA-2X3—complexes of mRNA with PLNs stabilized by 2X3; PHA-2X3-DOPE 1:1—complexes of mRNA with PLNs stabilized by a mixture of 2X3 and DOPE in a molar ratio of 1:1; PHA-2X3-DOPE 1:2—complexes of mRNA with PLNs stabilized with a mixture of 2X3 and DOPE in a molar ratio of 1:2. PHA-2X3-DOPE 1:3—complexes of mRNA with PLNs stabilized with a mixture of 2X3 and DOPE in a molar ratio of 1:3. PHA/lipids ratio in all the PLNs was 20:1 (wt.). Green line—signal of the fluorescent dye (used as an internal control for the electrophoresis). N/P — molar ratio between positively charged cationic liposomes and negatively charged delivered mRNA molecules.</p>
Full article ">Figure 3
<p>The efficiency of mRNA-eGFP delivery using CLs and PLNs in BHK-21 cells was measured using flow cytometry. The transfected cells were analyzed for the percentage of cells with detectable eGFP signals, and the MFI (mean fluorescence intensity) was recorded. Lipofectamine MessengerMAX (Lipofectamine MM) is a commercial transfection reagent and was used as a positive control for mRNA transfection. All the measurements were triplicated. The statistical analysis was performed using a two-way ANOVA: **—<span class="html-italic">p</span> &lt; 0.01; not significant—‘ns’.</p>
Full article ">Figure 4
<p>Dependences of the average particle diameter (d, nm), polydispersity (PDI), and ζ-potential (ζ, mV) on the N/P. Complexes of the PLNs (PHA-2X3-DOPE 1:3) and mRNA-eGFP were analyzed. PLNs (PHA-2X3-DOPE 1:3) were used as a control.</p>
Full article ">Figure 5
<p>(<b>a</b>) Efficiency of mRNA-Cy5 uptake with PLNs and CLs by BHK-21 cells determined by flow cytometry. Transfected cells—the percentage of fluorescent cells (Cy5 signal); MFI—mean fluorescence intensity. Statistical analysis was performed using one-way ANOVA: ***—<span class="html-italic">p</span> &lt; 0.001; **—<span class="html-italic">p</span> &lt; 0.01. (<b>b</b>) Fluorescence microscopy of cells transfected with mRNA-Cy5 mRNA complexes using PHA-2X3-DOPE 1:3, 2X3-DOPE 1:3, and Lipofectamine Messenger MAX (Lipofectamine MM); Scale bar 50 μm; cell nuclei–blue, eGFP—green, and mRNA–Cy5—red.</p>
Full article ">Figure 6
<p>Dependences of the average particle diameter (d, nm), polydispersity (PDI) and ζ-potential (ζ, mV) on the molar concentration of Na<sup>+</sup> in PBS. PLNs (PHA-2X3-DOPE 1:3) or CLs (2X3-DOPE 1:3) were mixed with PBS and incubated for either 2 h at 25 °C or 5 days at 4 °C.</p>
Full article ">Figure 7
<p>Transmission electron microscopy of PLNs (PHA-2X3-DOPE 1:3) and CLs (2X3-DOPE 1:3) at various Na<sup>+</sup> concentrations (0, 30, and 150 mM; incubation time 2 h). Scale bar: 150 nm (<b>A,B</b>), 100 nm (<b>C,E</b>), 200 nm (<b>D</b>), 500 nm (<b>F</b>).</p>
Full article ">
16 pages, 10397 KiB  
Article
Stiff-Soft Hybrid Biomimetic Nano-Emulsion for Targeted Liver Delivery and Treatment of Early Nonalcoholic Fatty Liver Disease
by Juan Li, Mingxing Yin, Maoxian Tian, Jianguo Fang and Hanlin Xu
Pharmaceutics 2024, 16(10), 1303; https://doi.org/10.3390/pharmaceutics16101303 - 7 Oct 2024
Viewed by 548
Abstract
Background: Nonalcoholic fatty liver disease (NAFLD) poses a risk for numerous metabolic diseases. To date, the U.S. Food and Drug Administration has not yet approved any medications for the treatment of NAFLD, for which developing therapeutic drugs is urgent. Dihydromyricetin (DMY), the most [...] Read more.
Background: Nonalcoholic fatty liver disease (NAFLD) poses a risk for numerous metabolic diseases. To date, the U.S. Food and Drug Administration has not yet approved any medications for the treatment of NAFLD, for which developing therapeutic drugs is urgent. Dihydromyricetin (DMY), the most abundant flavonoid in vine tea, has been shown to be hepatoprotective. Its application was limited by low bioavailability in vivo; Methods: In order to improve the bioavailability of DMY and achieve liver-targeted delivery, we designed a DMY-loaded stiff-soft hybrid biomimetic nano drug delivery system (DMY-hNE). The in vivo absorption, distribution, pharmacokinetic profiles, and anti-NAFLD efficacy of DMY-hNE were studied; Results: DMY-hNE was composed of a stiff core and soft shell, which led to enhanced uptake by gastrointestinal epithelial cells and increased penetration of the mucus barrier, thus improving the in vivo absorption, plasma DMY concentration, and liver distribution versus free DMY. In an early NAFLD mouse model, DMY-hNE effectively ameliorated fatty lesions accompanied with reduced lipid levels and liver tissue inflammation; Conclusions: These findings suggested that DMY-hNE is a promising platform for liver drug delivery and treatment of hepatopathy. Full article
(This article belongs to the Special Issue Delivery System for Biomacromolecule Drugs: Design and Application)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of in vivo absorption and liver-targeted delivery of DMY-loaded stiff-soft hybrid nano-emulsion (DMY-hNE).</p>
Full article ">Figure 2
<p>Characterization of DMY-hNE. (<b>A</b>) Size measured by DLS, (<b>B</b>) representative FTEM image and (<b>C</b>) STEM image of DMY-hNE in water. Scale bars represent 500 nm (black) and 200 nm (red).</p>
Full article ">Figure 3
<p>(<b>A</b>) Stability and (<b>B</b>) in vitro release profiles of DMY-hNE.</p>
Full article ">Figure 4
<p>(<b>A</b>) Stomach uptake of healthy mice i.g. treated with C6-hNE, C6 + vehicle, free C6 at 0.5 h. (<b>B</b>) Stomach uptake of healthy mice i.g. treated with C6-hNE at different time intervals (0.5, 1, 2 h). Scale bars represent 50 μm (white) and 20 μm (yellow).</p>
Full article ">Figure 5
<p>(<b>A</b>) Small intestine uptake of healthy mice i.g. treated with C6-hNE, C6 + vehicle, free C6 at 0.5 h. (<b>B</b>) Small intestine uptake of healthy mice i.g. treated with C6-hNE at different time intervals (0.5, 1, 2 h). Scale bars represent 50 μm (white) and 20 μm (yellow).</p>
Full article ">Figure 6
<p>Biodistribution and liver-targeting effect of DiR-hNE. (<b>A</b>) Representative ex vivo images and (<b>B</b>) quantitative analysis results in different time intervals. * <span class="html-italic">p</span> &lt; 0.05, vs. free DiR.</p>
Full article ">Figure 7
<p>In vivo pharmacokinetic profiles of DMY-hNE. DMY-hNE and free DMY were orally administered to rats at a dose of 25 mg DMY/kg, respectively (n = 3).</p>
Full article ">Figure 8
<p>In vivo anti-NAFLD efficacy of DMY-hNE. Free DMY, DMY + vehicle, and DMY-hNE were orally administrated to NAFLD mice at a dose of 25 mg DMY/kg, respectively, while MET was orally administrated at a dose of 250 mg MET/kg. (<b>A</b>) Schematic illustration of the experiment, (<b>B</b>) body weight changes, (<b>C</b>) liver index (Black dashed lines represent liver index range of healthy mice), and (<b>D</b>) representative HE and oil red O staining of healthy mice and NAFLD mice treated with water (placebo), free DMY, DMY + vehicle, DMY-hNE and metformin for 6 weeks. Scale bars represent 100 μm (black), 25 μm (yellow) and 200 μm (blue). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, vs. group Model.</p>
Full article ">Figure 9
<p>Blood biochemistry. Plasma (<b>A</b>) TG, (<b>B</b>) TC, (<b>C</b>) LDL, (<b>D</b>) HDL, (<b>E</b>) ALT, (<b>F</b>) AST, (<b>G</b>) ALP, (<b>H</b>) BUN, and (<b>I</b>) MDA levels of healthy mice and NAFLD mice treated with water (placebo), free DMY, DMY + vehicle, DMY-hNE and metformin for 6 weeks. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, vs. group Model.</p>
Full article ">Figure 10
<p>Liver tissue biochemistry. Liver (<b>A</b>) NO, (<b>B</b>) MDA, (<b>C</b>) GSH, (<b>D</b>) SOD, and (<b>E</b>) H<sub>2</sub>O<sub>2</sub> levels of healthy mice and NAFLD mice treated with water (placebo), free DMY, DMY + vehicle, DMY-hNE and metformin for 6 weeks. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, vs. group Model.</p>
Full article ">Figure 11
<p>In vivo safety evaluation of DMY-hNE. (<b>A</b>) Plasma ALT, AST, BUN levels and (<b>B</b>) representative HE staining of the main organs of healthy mice treated with water, free DMY, DMY + vehicle, DMY-hNE and metformin for 6 weeks. Scale bars represent 100 μm.</p>
Full article ">
60 pages, 1160 KiB  
Review
Synthesis, Photocatalytic and Bio Activity of ZnO-TiO2 Nanocomposites: A Review Study
by Fulvia Pinzari
Reactions 2024, 5(4), 680-739; https://doi.org/10.3390/reactions5040035 - 2 Oct 2024
Viewed by 377
Abstract
Zinc oxide and titanium dioxide are materials with strong photocatalytic and antimicrobial activity. This activity is greater when the material is in nanocrystalline form. It has been seen that these properties are also present in the ZnO-TiO2 nanocomposite material, and the extent [...] Read more.
Zinc oxide and titanium dioxide are materials with strong photocatalytic and antimicrobial activity. This activity is greater when the material is in nanocrystalline form. It has been seen that these properties are also present in the ZnO-TiO2 nanocomposite material, and the extent depends on multiple factors, such as crystallinity, structural composition, crystallite size, and morphology. These structural properties can be varied by acting on the synthesis of the material, obtaining a wide variety of composites: random nanoparticles, nanorods, nanowires, nanotubes, nanofibers, tetrapods, core–shell, hollow spheres, inverse opal structures (IOSs), hierarchical structures, and films. When an interface between nanocrystallites of the two oxides is created, the composite system manages to have photocatalytic activity greater than that of the two separate oxides, and in certain circumstances, even greater than P25. The antimicrobial activity results also improved for the composite system compared to the two separate oxides. These two aspects make these materials interesting in various fields, such as wastewater and air treatment, energy devices, solar filters, and pharmaceutical products and in the context of the restoration of monumental cultural assets, in which their use has a preventive purpose in the formation of biofilms. In this review we analyse the synthesis techniques of ZnO-TiO2 nanocomposites, correlating them to the shape obtained, as well as the photocatalytic and antimicrobial activity. It is also illustrated how ZnO-TiO2 nanocomposites can have a less negative impact on toxicity for humans and the environment compared to the more toxic ZnO nanoparticles or ZnO. Full article
(This article belongs to the Special Issue Nanoparticles: Synthesis, Properties, and Applications)
Show Figures

Scheme 1

Scheme 1
<p>Different ZnO-TiO<sub>2</sub> composite shapes synthetized.</p>
Full article ">Scheme 2
<p>Alternative synthesis methodologies used for the simpler ZTNC shapes.</p>
Full article ">Scheme 3
<p>Sub-shapes of the more complex shapes.</p>
Full article ">Scheme 4
<p>Factors influencing the increased photocatalytic effect of ZTNCs compared to TNPs and ZNPs.</p>
Full article ">
24 pages, 1486 KiB  
Article
Finite Nuclear Size Effect on the Relativistic Hyperfine Splittings of 2s and 2p Excited States of Hydrogen-like Atoms
by Katharina Lorena Franzke and Uwe Gerstmann
Foundations 2024, 4(4), 513-536; https://doi.org/10.3390/foundations4040034 - 1 Oct 2024
Viewed by 414
Abstract
Hyperfine splittings play an important role in quantum information and spintronics applications. They allow for the readout of the spin qubits, while at the same time providing the dominant mechanism for the detrimental spin decoherence. Their exact knowledge is thus of prior relevance. [...] Read more.
Hyperfine splittings play an important role in quantum information and spintronics applications. They allow for the readout of the spin qubits, while at the same time providing the dominant mechanism for the detrimental spin decoherence. Their exact knowledge is thus of prior relevance. In this work, we analytically investigate the relativistic effects on the hyperfine splittings of hydrogen-like atoms, including finite-size effects of the nucleis’ structure. We start from exact solutions of Dirac’s equation using different nuclear models, where the nucleus is approximated by (i) a point charge (Coulomb potential), (ii) a homogeneously charged full sphere, and (iii) a homogeneously charged spherical shell. Equivalent modelling has been done for the distribution of the nuclear magnetic moment. For the 1s ground state and 2s excited state of the one-electron systems H1, H2, H3, and He+3, the calculated finite-size related hyperfine shifts are quite similar for the different structure models and in excellent agreement with those estimated by comparing QED and experiment. This holds also in a simplified approach where relativistic wave functions from a Coulomb potential combined with spherical-shell distributed nuclear magnetic moments promises an improved treatment without the need for an explicit solution of Dirac’s equation within the nuclear core. Larger differences between different nuclear structure models are found in the case of the anisotropic 2p3/2 orbitals of hydrogen, rendering these excited states as promising reference systems for exploring the proton structure. Full article
(This article belongs to the Section Physical Sciences)
Show Figures

Figure 1

Figure 1
<p><b>Left:</b> Potential <math display="inline"><semantics> <mrow> <mi>V</mi> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> for <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>1</mn> </mmultiscripts> </semantics></math> generated by a point charge <math display="inline"><semantics> <mrow> <mi>Z</mi> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math> and by homogeneously charged full spheres and spherical shells, both with <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mi mathvariant="normal">H</mi> </msub> <mo>=</mo> <mn>0.83</mn> </mrow> </semantics></math> fm (radius taken from [<a href="#B24-foundations-04-00034" class="html-bibr">24</a>]). All three potentials match outside the nucleus, i.e., for <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>≥</mo> <msub> <mi>R</mi> <mi mathvariant="normal">H</mi> </msub> </mrow> </semantics></math>. <b>Right:</b> Radial wave function <math display="inline"><semantics> <mrow> <mi>g</mi> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> for the <math display="inline"><semantics> <mrow> <mn>1</mn> <mi>s</mi> </mrow> </semantics></math> state of the three hydrogen isotopes for the three potentials. Note that the difference due to the different nuclear charge models is similar to that caused by the different isotopes (isotope shift).</p>
Full article ">Figure 2
<p><math display="inline"><semantics> <mrow> <mi>S</mi> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> (<b>top</b>) for hydrogen <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>1</mn> </mmultiscripts> </semantics></math>, deuterium <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>2</mn> </mmultiscripts> </semantics></math>, and tritium <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>3</mn> </mmultiscripts> </semantics></math> for all three nuclei models. <math display="inline"><semantics> <mrow> <mi>S</mi> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> </semantics></math> matches for all three potentials if <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>≥</mo> <mi>R</mi> </mrow> </semantics></math>. The inset shows the non-relativistic case (<math display="inline"><semantics> <mrow> <mi>S</mi> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> <mo>=</mo> <mn>1</mn> </mrow> </semantics></math>) together with the relativistic form function of a pure Coulomb potential. The derivative <math display="inline"><semantics> <mrow> <msup> <mi>S</mi> <mo>′</mo> </msup> <mrow> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> <mo>=</mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mi>d</mi> <mi>s</mi> </mrow> <mrow> <mi>d</mi> <mi>r</mi> </mrow> </mfrac> </mstyle> </mrow> </semantics></math> of the relativistic form function is shown for hydrogen, deuterium, and tritium for all three nuclei models (<b>bottom</b>); <math display="inline"><semantics> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mrow> <mo>∂</mo> <mi>S</mi> </mrow> <mrow> <mo>∂</mo> <mi>r</mi> </mrow> </mfrac> </mstyle> </semantics></math> matches for all three potentials if <math display="inline"><semantics> <mrow> <mi>r</mi> <mo>≥</mo> <mi>R</mi> </mrow> </semantics></math>.</p>
Full article ">Figure 3
<p>Electron probability <math display="inline"><semantics> <mrow> <msup> <mi>ψ</mi> <mn>2</mn> </msup> <mrow> <mo stretchy="false">(</mo> <mi>r</mi> <mo stretchy="false">)</mo> </mrow> <mo>·</mo> <msup> <mi>r</mi> <mn>2</mn> </msup> </mrow> </semantics></math> (<span class="html-italic">r</span> in fm) of the excited <math display="inline"><semantics> <mrow> <mn>2</mn> <msub> <mi>p</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> state for the hydrogen isotopes <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>1</mn> </mmultiscripts> </semantics></math>, <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>2</mn> </mmultiscripts> </semantics></math>, <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>3</mn> </mmultiscripts> </semantics></math>, and <math display="inline"><semantics> <mmultiscripts> <mi>He</mi> <none/> <mo>+</mo> <mprescripts/> <none/> <mn>3</mn> </mmultiscripts> </semantics></math> (curve for the latter multiplied by 500). For <math display="inline"><semantics> <mmultiscripts> <mi mathvariant="normal">H</mi> <none/> <none/> <mprescripts/> <none/> <mn>1</mn> </mmultiscripts> </semantics></math>, beside that for the full-sphere nucleus (solid lines) also the curve for the spherical-shell model (dashed line) is shown.</p>
Full article ">
13 pages, 2203 KiB  
Article
Synthesis, Material Properties, and Organocatalytic Performance of Hypervalent Iodine(III)-Oxidants in Core–Shell-Structured Magnetic Nanoparticles
by Julien Grand, Carole Alayrac, Simona Moldovan and Bernhard Witulski
Catalysts 2024, 14(10), 677; https://doi.org/10.3390/catal14100677 - 1 Oct 2024
Viewed by 395
Abstract
Magnetic nanoparticles (MNPs) based on magnetite (Fe3O4) are attractive catalyst supports due to their high surface area, easy preparation, and facile separation, but they lack stability in acidic reaction media. The search for MNPs stable in oxidative acidic reaction [...] Read more.
Magnetic nanoparticles (MNPs) based on magnetite (Fe3O4) are attractive catalyst supports due to their high surface area, easy preparation, and facile separation, but they lack stability in acidic reaction media. The search for MNPs stable in oxidative acidic reaction media is a necessity if one wants to combine the advantages of MNPs as catalyst supports with those of iodine(III) reagents being environmentally benign oxidizers. In this work, immobilized iodophenyl organocatalysts on magnetite support (IMNPs) were obtained by crossed-linking polymerization of 4-iodostyrene with 1,4-divinylbenzene in the presence of MNPs. The obtained IMNPs were characterized by TGA, IR, SEM, STEM, and HAADF to gain information on catalyst morphology, average particle size (80–100 nm), and their core–shell structure. IMNP-catalysts tested in (i) the α-tosyloxylation of propiophenone 1 with meta-chloroperbenzoic acid (m-CPBA) and (ii) in the oxidation of 9,10-dimethoxyanthracene 3 with Oxone® as the side-oxidant showed a similar performance as reactions using stoichiometric amounts of iodophenyl. The developed IMNPs withstand strong acidic conditions and serve as reusable organocatalysts. They are recyclable up to four times for repeated organocatalytic oxidations with rates of recovery of 80–92%. This is the first example of a—(4-iodophenyl)polystyrene shell—magnetite core-structured organocatalyst withstanding strong acidic reaction conditions. Full article
Show Figures

Figure 1

Figure 1
<p>Reaction mixture containing (<b>a</b>) dispersed IMNPs in the absence of an external magnet and (<b>b</b>) IMNPs collected using an external permanent magnet.</p>
Full article ">Figure 2
<p>(<b>A</b>) Powder XRD pattern of (a) magnetite nanoparticles (MNPs) and (b) iodinated core-shell-structured magnetic nanoparticles (IMNPs); (<b>B</b>) ATR-FT-IR spectra of (a) MNPs, (b) HDA@MNPs, and (c) IMNPs.</p>
Full article ">Figure 3
<p>SEM images of obtained (<b>A</b>) magnetite MNPs and (<b>B</b>) obtained IMNPs (see also <a href="#app1-catalysts-14-00677" class="html-app">Supplementary Materials</a>) together with their particle size distribution histogram for 55 measured particles of (<b>C</b>) MNPs and (<b>D</b>) IMNPs, as determined from the SEM image.</p>
Full article ">Figure 4
<p>STEM-HAADF micrographs of obtained magnetite MNPs (upper row) and those of IMNPs. (<b>a</b>) General view of MNP agglomerations with the size distribution in inset; (<b>b</b>) HR-STEM-HAADF image of several MNPs and the FFT in the inset confirm the crystal structure of Fe<sub>3</sub>O<sub>4</sub>; (<b>c</b>) spatial distribution of Fe<sub>3</sub>O<sub>4</sub> IMNPs on the polymer granules and statistical size distribution of the polymer; and (<b>d</b>) the “idealized core shell structure” of IMNPs.</p>
Full article ">Scheme 1
<p>Preparation of Fe<sub>3</sub>O<sub>4</sub>-based magnetic nanoparticles stabilized with 16-heptadecenoic acid (HDA@MNP) and their conversion to “core–shell-structured” iodinated magnetic nanoparticles (IMNPs).</p>
Full article ">Scheme 2
<p>Organocatalytic α-tosyloxylation of propiophenone (<b>1</b>) with IMNPs and <span class="html-italic">m</span>-CPBA as the side oxidant to give α-tosyloxypropiophenone <b>2</b>.</p>
Full article ">Scheme 3
<p>Organocatalytic oxidation of 9,10-dimethoxyanthracene (<b>3</b>) with sub-stoichiometric amounts of IMNPs (30 mol%) and with Oxone<sup>®</sup> as the side-oxidant to give anthraquinone <b>4</b>.</p>
Full article ">
12 pages, 2485 KiB  
Article
Spectral Control by Silver Nanoparticle-Based Metasurfaces for Mitigation of UV Degradation in Perovskite Solar Cells
by Silvia Delgado-Rodríguez, Eva Jaldo Serrano, Mahmoud H. Elshorbagy, Javier Alda, Gonzalo del Pozo and Alexander Cuadrado
Nanomaterials 2024, 14(19), 1582; https://doi.org/10.3390/nano14191582 - 30 Sep 2024
Viewed by 420
Abstract
Perovskite solar cells are considered to be one of the most promising solar cell designs in terms of photovoltaic efficiency. However, their practical deployment is strongly affected by their short lifetimes, mostly caused by environmental conditions and UV degradation. In this contribution, we [...] Read more.
Perovskite solar cells are considered to be one of the most promising solar cell designs in terms of photovoltaic efficiency. However, their practical deployment is strongly affected by their short lifetimes, mostly caused by environmental conditions and UV degradation. In this contribution, we present a metasurface made of silver nanoparticles used as a UV filter on a perovskite solar cell. The UV-blocking layer was fabricated and morphologically and compositionally analyzed. Its optical response, in terms of optical transmission, was also experimentally measured. These results were compared with simulations made through the use of a well-proven computational electromagnetism model. After analyzing the discrepancies between the experimental and simulated results and checking those obtained from electron beam microscopy and electron dispersion spectroscopy, we could see that a residue from fabrication, sodium citrate, strongly modified the optical response of the system, generating a redshift of about 50 nm. Then, we proposed and simulated the optical behavior of core–shell nanoparticles made of silver and silica. The calculated spectral absorption at the active perovskite layer shows how the appropriate selection of the geometrical parameters of these core–shell particles is able to tune the absorption at the active layer by removing a significant portion of the UV band and reducing the absorption of the active layer from 90% to 5% at a resonance wavelength of 403 nm. Full article
Show Figures

Figure 1

Figure 1
<p>Calculation of the absorption coefficients <span class="html-italic">Q<sub>ext</sub></span>, <span class="html-italic">Q<sub>abs</sub></span>, and <span class="html-italic">Q<sub>sc</sub></span>, and the electric field distributions for several cases involving Ag NP. We show the simulation results for (<b>a</b>) a 20 nm Ag NPs surrounded by air, (<b>c</b>) a 20 nm Ag NPs surrounded by air on a glass substrate (<b>e</b>) an 80 nm Ag NPs surrounded by air on glass substrate. (<b>b</b>,<b>d</b>,<b>f</b>) are the normalized electric field distributions for the cases in (<b>a</b>,<b>c</b>,<b>e</b>), respectively.</p>
Full article ">Figure 2
<p>(<b>Top</b>): Electron microscope photography of the sample showing the presence of silver nanoparticles on the glass substrate, and citrate residue (boxes). The NPs are embedded within a medium. (<b>Bottom</b>): EDS results for the regions marked in the top photography. This spectrogram shows the presence of Na, O<sub>2</sub>, and C, that are part of sodium citrate.</p>
Full article ">Figure 3
<p>(<b>a</b>) Spectral transmittance obtained experimentally from the samples (in dashed black line), and simulated for a periodic arrangement with period 122 nm for a collection of 20 nm silver spherical NPs located on a glass substrate (in dashed green) and for the same collection of NPs surrounded by a spherical cell of thickness 20 nm of a dielectric material having an index of refraction n = 1.58 (equivalent to the sodium citrate spherical shell). (<b>b</b>,<b>c</b>) Maps of the electric field modulus for the case of the equivalent sodium citrate spherical shell, and the silver NP surrounded by air, respectively. Both cases have the NPs located on a BK7 glass substrate.</p>
Full article ">Figure 4
<p>(<b>a</b>) Simulated optical transmission of the UV filter based on Ag nanoparticle. (<b>b</b>) Simulated spectral absorptance at the filtered perovskite active layer for the cases of interest. The grey dotted line is for the solar cell alone. To both figures, the blue line is for a regular arrangement of silver NPs, 20 nm in diameter and spatial period of 40 nm, while the red dashed line is for the same NPs but coated with a dielectric layer which simulates the sodium citrate around the silver NPs.</p>
Full article ">Figure 5
<p>(<b>a</b>) Simulated optical transmission of the UV filter based on silver–silica core–shell nanoparticle. (<b>b</b>) Simulated spectral absorption at the active perovskite layer showing the result for the regular cell (grey dotted line). To both figures, the blue line is for a periodic arrangement of core–shell silver-silica NPs, and the red dashed line is related to a periodic arrangement is coated with a material that reproduces the results of sodium citrate residue.</p>
Full article ">
14 pages, 2923 KiB  
Article
Facile Synthesis of Core-Shell Magnetic Iron Oxide@SiO2-NH2 Nanoparticles and Their Application in Rapid Boron Removal from Aqueous Solutions
by Qinqin Hu, Manman Zhang, Jiaoyu Peng, Yaping Dong, Wu Li and Lingzong Meng
Magnetochemistry 2024, 10(10), 74; https://doi.org/10.3390/magnetochemistry10100074 - 30 Sep 2024
Viewed by 385
Abstract
In this study, amino-functionalized magnetic particles (iron oxide@SiO2-NH2) with core-shell structures were synthesized and evaluated for rapid boron removal from aqueous solutions. The results showed that the specific surface area of the iron oxide@SiO2-NH2 (131.24 m [...] Read more.
In this study, amino-functionalized magnetic particles (iron oxide@SiO2-NH2) with core-shell structures were synthesized and evaluated for rapid boron removal from aqueous solutions. The results showed that the specific surface area of the iron oxide@SiO2-NH2 (131.24 m2⋅g−1) increased greatly compared to pure iron oxide (30.98 m2⋅g−1). The adsorption equilibrium was less than 2 h, with an adsorption capacity of 29.76 mg⋅g−1 at pH = 6 at 15 °C. The quasi-second-order kinetic model described the boron adsorption process well, and both the Langmuir and Freundlich models were suitable for characterizing the adsorption isotherms. The zeta potential and XPS analysis before and after adsorption revealed that the main adsorption mechanism was the hydrogen bonding formation between the terminal -NH2 groups of the adsorbent and the boric acid. In addition, the adsorbent still maintained a high adsorption performance after five adsorption–desorption cycles, which illustrated that the iron oxide@SiO2-NH2 may be a potential adsorbent for environmental boron removal treatment. Full article
(This article belongs to the Section Magnetic Nanospecies)
Show Figures

Figure 1

Figure 1
<p>Effects of (<b>a</b>) APTES and (<b>b</b>) TEOS dosage on the synthesis of iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern and (<b>b</b>) FT-IR spectrum of the iron oxide, iron oxide@SiO<sub>2</sub>, and iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 3
<p>SEM (<b>a</b>,<b>b</b>) and TEM (<b>a<sub>i</sub></b>,<b>b<sub>i</sub></b>) images of the iron oxide and the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 4
<p>EDS spectra of the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isotherms, pore size distribution curves (the inset), and (<b>b</b>) thermogravimetric analysis curves of the iron oxide and the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Hysteresis loops and (<b>b</b>) pH dependent on the zeta potential of the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 7
<p>Effect of different factors on the adsorption capacity: (<b>a</b>) dosage; (<b>b</b>) solution pH; (<b>c</b>) initial boron concentration, and (<b>d</b>) contact time.</p>
Full article ">Figure 8
<p>XPS spectra of (<b>a</b>) the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>, (<b>b</b>) B 1s, and (<b>c</b>) N 1s before and after boron adsorption.</p>
Full article ">Figure 9
<p>Plots of (<b>a</b>) pseudo-first-order kinetic model and (<b>b</b>) pseudo-second-order kinetic model of boron adsorption on the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Isotherms of boron adsorption on iron oxide@SiO<sub>2</sub>-NH<sub>2</sub>; fitting curves of (<b>b</b>) Langmuir model and (<b>c</b>) Freundlich model.</p>
Full article ">Figure 11
<p>Adsorption–desorption cycles of the iron oxide@SiO<sub>2</sub>–NH<sub>2</sub> about boron adsorption.</p>
Full article ">Scheme 1
<p>Schematic preparation of the iron oxide@SiO<sub>2</sub>-NH<sub>2</sub> MNPs.</p>
Full article ">
10 pages, 4125 KiB  
Article
Preparation and Properties of Thermoregulated Seaweed Fibers Based on Magnetic Paraffin wax@calcium Carbonate Microcapsules
by Yonggui Li, Congzhu Xu, Yuanxin Lin, Xiaolei Song, Runjun Sun, Qiang Wang and Xinqun Feng
Materials 2024, 17(19), 4826; https://doi.org/10.3390/ma17194826 - 30 Sep 2024
Viewed by 361
Abstract
In order to enhance the application of thermoregulated materials, magnetic phase change microcapsules were prepared using a self-assembly method. Paraffin wax was chosen for its fine thermoregulation properties as the core material, while Fe3O4 nanoparticles doped in calcium carbonate served [...] Read more.
In order to enhance the application of thermoregulated materials, magnetic phase change microcapsules were prepared using a self-assembly method. Paraffin wax was chosen for its fine thermoregulation properties as the core material, while Fe3O4 nanoparticles doped in calcium carbonate served as the hybrid shell material. The microcapsules were then blended with sodium alginate and processed into seaweed fibers through wet spinning. The microstructure, thermal, and magnetic properties of the microcapsules were analyzed using scanning electron microscopy, energy dispersive X-ray spectroscopy, a laser particle size analyzer, an X-ray diffractometer, a differential scanning calorimeter, a thermogravimetric analyzer, and a vibrating sample magnetometer. The thermoregulation of the fibers was evaluated using a thermal infrared imager. The results indicated that the microcapsules had a uniform size distribution and good thermal properties. When the mass fraction of Fe3O4 nanoparticles was 8%, the microcapsules exhibited a saturation magnetization of 2.44 emu/g and an enthalpy value of 94.25 J/g, indicating effective phase change and magnetic properties. Furthermore, the thermoregulated seaweed fibers showed a high enthalpy value of 19.8 J/g with fine shape, offering potential for developing multifunctional fiber products. Full article
(This article belongs to the Special Issue Synthesis and Properties of Flame Retardant for Polymers)
Show Figures

Figure 1

Figure 1
<p>Mechanism of preparation of thermoregulated seaweed fibers.</p>
Full article ">Figure 2
<p>(<b>a</b>,<b>b</b>) SEM images of PW@CaCO<sub>3</sub> phase change microcapsules; (<b>c</b>,<b>d</b>) SEM image of PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) Particle size distribution images of PW@CaCO<sub>3</sub> phase change microcapsules and PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules; (<b>c</b>,<b>d</b>) EDS image of PW@CaCO<sub>3</sub> phase change microcapsules and PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD patterns of CaCO<sub>3</sub>, Fe<sub>3</sub>O<sub>4</sub> nanoparticles, and PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules; (<b>b</b>) DSC curves of PW and microcapsules; (<b>c</b>) TG curves of PW and microcapsules; (<b>d</b>) VSM curves of magnetic phase change microcapsules.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Physical drawings of PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules and seaweed fibers; (<b>c</b>,<b>d</b>) SEM of seaweed fibers with PW@CaCO<sub>3</sub>@8% Fe<sub>3</sub>O<sub>4</sub> phase change microcapsules at different magnifications.</p>
Full article ">Figure 6
<p>(<b>a</b>) Heating and (<b>b</b>) cooling DSC curves in seaweed fibers.</p>
Full article ">Figure 7
<p>(<b>a</b>) Heating curves and (<b>b</b>) cooling curves in time-temperature graph of seaweed fibers.</p>
Full article ">
Back to TopTop