[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
remove_circle_outline
remove_circle_outline

Search Results (407)

Search Parameters:
Keywords = silica nanocomposite

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 8179 KiB  
Article
Unraveling the Effect of Strain Rate and Temperature on the Heterogeneous Mechanical Behavior of Polymer Nanocomposites via Atomistic Simulations and Continuum Models
by Ali A. Youssef, Hilal Reda and Vagelis Harmandaris
Polymers 2024, 16(17), 2530; https://doi.org/10.3390/polym16172530 - 6 Sep 2024
Viewed by 379
Abstract
Polymer nanocomposites are characterized by heterogeneous mechanical behavior and performance, which is mainly controlled by the interaction between the nanofiller and the polymer matrix. Optimizing their material performance in engineering applications requires understanding how both the temperature and strain rate of the applied [...] Read more.
Polymer nanocomposites are characterized by heterogeneous mechanical behavior and performance, which is mainly controlled by the interaction between the nanofiller and the polymer matrix. Optimizing their material performance in engineering applications requires understanding how both the temperature and strain rate of the applied deformation affect mechanical properties. This work investigates the effect of strain rate and temperature on the mechanical properties of poly(ethylene oxide)/silica (PEO/SiO2) nanocomposites, revealing their behavior in both the melt and glassy states, via atomistic molecular dynamics simulations and continuum models. In the glassy state, the results indicate that Young’s modulus increases by up to 99.7% as the strain rate rises from 1.0 × 10−7 fs−1 to 1.0 × 10−4 fs−1, while Poisson’s ratio decreases by up to 39.8% over the same range. These effects become even more pronounced in the melt state. Conversely, higher temperatures lead to an opposing trend. A local, per-atom analysis of stress and strain fields reveals broader variability in the local strain of the PEO/SiO2 nanocomposites as temperature increases and/or the deformation rate decreases. Both interphase and matrix regions lose rigidity at higher temperatures and lower strain rates, blurring their distinctiveness. The results of the atomistic simulations concerning the elastic modulus and Poisson’s ratio are in good agreement with the predictions of the Richeton–Ji model. Additionally, these findings can be leveraged to design advanced polymer composites with tailored mechanical properties and could optimize structural components by enhancing their performance under diverse engineering conditions. Full article
(This article belongs to the Special Issue Rheological Properties of Polymers and Polymer Composites)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Snapshots of the PEO/SiO<sub>2</sub> model systems, using periodically wrapped coordinates, in (<b>a</b>) initial (equilibrium) and (<b>b</b>) deformed in the x direction (ε = 0.4) configurations at T = 270 K. Blue dots represent the nearly spherical SiO<sub>2</sub> nanoparticle, while red dots represent the surrounding PEO polymer in the unit cell system.</p>
Full article ">Figure 2
<p>Variations in mechanical properties with temperature and strain rate: (<b>a</b>) Young modulus, (<b>b</b>) Poisson ratio. Darker colors indicate a transition towards a glassy state (higher E &amp; lower ν values), while lighter colors represent a more fluid, melty state (lower E &amp; higher ν values).</p>
Full article ">Figure 3
<p>Linear mechanical properties of the PEO/SiO<sub>2</sub> nanocomposites model as a function of strain rate: (<b>a</b>) Young’s modulus, (<b>b</b>) Poisson’s ratio. The error bars are computed by analyzing ten uncorrelated configurations. The percentage error values between the MD simulation results and the RJ model are displayed above each data point on the plot.</p>
Full article ">Figure 4
<p>Linear mechanical properties of the PEO/SiO<sub>2</sub> nanocomposite model as a function of temperature: (<b>a</b>) Young’s modulus, (<b>b</b>) Poison’s ratio. The error bars are computed by analyzing ten uncorrelated configurations. The percentage error values between the MD simulation results and the RJ model are displayed above each data point on the plot.</p>
Full article ">Figure 5
<p>Schematic overview of designated zones within spheres around the nanoparticle in an in-plane box section.</p>
Full article ">Figure 6
<p>Elasticity modulus variations in (<b>a</b>) interphase and (<b>b</b>) matrix regions with temperature and strain rate dependency.</p>
Full article ">Figure 7
<p>PNC’s elasticity modulus as a function of strain rate for systems at temperatures (<b>a</b>) below <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math> (220 K), (<b>b</b>) equal to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math> (270 K), and (<b>c</b>) above <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math> (330 K).</p>
Full article ">Figure 8
<p>Elasticity modulus of PEO/SiO<sub>2</sub> nanocomposites as a function of temperature for systems at strain rates (<b>a</b>) 1.0 × 10<sup>−5</sup> fs<sup>−1</sup> and (<b>b</b>) 1.0 × 10<sup>−6</sup> fs<sup>−1</sup>.</p>
Full article ">Figure 9
<p>Analysis of the standard deviation with respect to the global strain and (inset) probability distribution of the local strain. The red, black, and blue curves represent the probability distribution in frames corresponding to global strains of 0.03, 0.06, and 0.09, respectively. The line styles for the red and blue inset plots match the legend of the black inset. (<b>a</b>–<b>c</b>) For bulk regions at 220 K, 270 K, and 330 K, respectively; and (<b>d</b>–<b>f</b>) for interphase regions at 220 K, 270 K, and 330 K, respectively.</p>
Full article ">Figure 10
<p>Variation in the mean-squared displacement (<b>a</b>) along x-direction and (<b>b</b>) for all components (sum of MSD along x, y, and z directions) in the interphase and matrix regions as a function of global applied strain under different strain rates, for systems at temperatures below (220 K), at (270 K), and above (330 K) the glass transition temperature (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure 10 Cont.
<p>Variation in the mean-squared displacement (<b>a</b>) along x-direction and (<b>b</b>) for all components (sum of MSD along x, y, and z directions) in the interphase and matrix regions as a function of global applied strain under different strain rates, for systems at temperatures below (220 K), at (270 K), and above (330 K) the glass transition temperature (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>T</mi> </mrow> <mrow> <mi>g</mi> </mrow> </msub> </mrow> </semantics></math>).</p>
Full article ">Figure A1
<p>Temperature effect (150–400 K) on systems at constant strain rates: (<b>a</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 1.0 × 10<sup>−5</sup> (<span class="html-italic">fs</span><sup>−1</sup>) and (<b>b</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 1.0 × 10<sup>−6</sup> (<span class="html-italic">fs</span><sup>−1</sup>).</p>
Full article ">Figure A2
<p>Strain rate effect on systems at constant temperatures: (<b>a</b>) T220 K, (<b>b</b>) T270 K, and (<b>c</b>) T330 K.</p>
Full article ">
16 pages, 5195 KiB  
Article
Polyurethane Nanocomposite Coatings Coupled with Titanium-Based Conversion Layers for Enhanced Anticorrosion, Icephobic Properties, and Surface Protection
by Shamim Roshan, Reza Jafari and Gelareh Momen
Molecules 2024, 29(16), 3901; https://doi.org/10.3390/molecules29163901 - 17 Aug 2024
Viewed by 470
Abstract
This study examines the efficacy of icephobic polyurethane nanocomposite coatings in mitigating corrosion on an aluminum substrate. A titanium-based conversion coating is applied to modify the substrate, and the research focuses on optimizing the dual functionalities of icephobicity and anticorrosion within the polyurethane [...] Read more.
This study examines the efficacy of icephobic polyurethane nanocomposite coatings in mitigating corrosion on an aluminum substrate. A titanium-based conversion coating is applied to modify the substrate, and the research focuses on optimizing the dual functionalities of icephobicity and anticorrosion within the polyurethane coatings while ensuring strong substrate adhesion. The coatings are formulated using fluoropolyol, isocyanate, and silica nanoparticles treated with polydimethylsiloxane. Surface properties are analyzed using contact angles, contact angle hysteresis measurements, and atomic force microscopy, and the coatings’ icephobicity is evaluated through differential scanning calorimetry, freezing time delay, ice adhesion under impact and non-impact conditions, and ice accretion tests. The corrosion resistance and adhesive strength of the coatings are assessed using electrochemical impedance spectroscopy and cross-cut tests, respectively. Increasing the concentration of silica nanoparticles to 10 wt.% increases contact angles to 167°, although the 4 wt.% coating produces the lowest contact angle hysteresis (3° ± 0.5°) and ice nucleation temperature (−23 °C). The latter coating is then applied to a substrate pretreated with a titanium/cerium-based conversion coating. This prepared surface maintains an ice adhesion of about 15 kPa after 15 icing/de-icing cycles and provides approximately 90 days of surface protection (|Z|lf = 1.6 × 109 Ω·cm2). Notably, the impedance value exceeds that of untreated substrates, underscoring the effectiveness of the titanium/cerium-based conversion coating in enhancing both corrosion resistance and coating adhesion to the substrate. Full article
Show Figures

Figure 1

Figure 1
<p>Atomic force microscopy (AFM) images of the coatings’ topography: (<b>a</b>) FPU, (<b>b</b>) FPU-SiO<sub>2</sub> 4%, (<b>c</b>) FPU-SiO<sub>2</sub> 7%, and (<b>d</b>) FPU-SiO<sub>2</sub> 10%.</p>
Full article ">Figure 2
<p>(<b>a</b>) Evaluation by differential scanning calorimetry (DSC) of ice nucleation temperatures; (<b>b</b>) freezing delay time for the FPU and FPU-SiO<sub>2</sub> 4% coatings.</p>
Full article ">Figure 3
<p>(<b>a</b>) Ice adhesion strength measurements determined using push-off and centrifuge tests for the FPU and FPU-SiO<sub>2</sub> 4% samples; (<b>b</b>) ice adhesion strength; (<b>c</b>) sliding angle of the FPU and FPU-SiO<sub>2</sub> 4% coatings over 15 icing/de-icing cycles; (<b>d</b>) optical images of the FPU and FPU-SiO<sub>2</sub> 4% samples before and after the centrifuge test; (<b>e</b>) infrared thermographs of the FPU and FPU-SiO<sub>2</sub> 4% samples during 20 min of precipitation at an incline of 80°; (<b>f</b>) optical images of the FPU and FPU-SiO<sub>2</sub> 4% samples after 20 min of precipitation at an incline of 80°.</p>
Full article ">Figure 4
<p>Nyquist and Bode diagrams of FPU-SiO<sub>2</sub> 4% and FPU-SiO<sub>2</sub> 4%-CC samples immersed in a NaCl 3.5 wt.% solution (<b>a</b>) after 1 day; (<b>b</b>) after 60 days; (<b>c</b>) after 90 days.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic of the proposed mechanism of corrosion protection of FPU-SiO<sub>2</sub> 4%-CC; (<b>b</b>) appearance of the FPU-SiO<sub>2</sub> 4% and FPU-SiO<sub>2</sub> 4%-CC coating surfaces after the cross-cut test.</p>
Full article ">Figure 6
<p>(<b>a</b>) Contact angle and (<b>b</b>) contact angle hysteresis of the FPU, FPU-SiO<sub>2</sub> 4%, and FPU-SiO<sub>2</sub> 4%-CC over 28 days of exposure during a weathering test.</p>
Full article ">
12 pages, 4010 KiB  
Article
Improving Shale Stability through the Utilization of Graphene Nanopowder and Modified Polymer-Based Silica Nanocomposite in Water-Based Drilling Fluids
by Yerlan Kanatovich Ospanov, Gulzhan Abdullaevna Kudaikulova, Murat Smanovich Moldabekov and Moldir Zhumabaevna Zhaksylykova
Processes 2024, 12(8), 1676; https://doi.org/10.3390/pr12081676 - 10 Aug 2024
Viewed by 665
Abstract
Shale formations present significant challenges to traditional drilling fluids due to fluid infiltration, cuttings dispersion, and shale swelling, which can destabilize the wellbore. While oil-based drilling fluids (OBM) effectively address these concerns about their environmental impact, their cost limits their widespread use. Recently, [...] Read more.
Shale formations present significant challenges to traditional drilling fluids due to fluid infiltration, cuttings dispersion, and shale swelling, which can destabilize the wellbore. While oil-based drilling fluids (OBM) effectively address these concerns about their environmental impact, their cost limits their widespread use. Recently, nanomaterials (NPs) have emerged as a promising approach in drilling fluid technology, offering an innovative solution to improve the efficiency of water-based drilling fluids (WBDFs) in shale operations. This study evaluates the potential of utilizing modified silica nanocomposite and graphene nanopowder to formulate a nanoparticle-enhanced water-based drilling fluid (NP-WBDF). The main objective is to investigate the impact of these nanoparticle additives on the flow characteristics, filtration efficiency, and inhibition properties of the NP-WBDF. In this research, a silica nanocomposite was successfully synthesized using emulsion polymerization and analyzed using FTIR, PSD, and TEM techniques. Results showed that the silica nanocomposite exhibited a unimodal particle size distribution ranging from 38 nm to 164 nm, with an average particle size of approximately 72 nm. Shale samples before and after interaction with the graphene nanopowder WBDF and the silica nanocomposite WBDF were analyzed using scanning electron microscopy (SEM). The NP-WBM underwent evaluation through API filtration tests (LTLP), high-temperature/high-pressure (HTHP) filtration tests, and rheological measurements conducted with a conventional viscometer. Experimental results showed that the silica nanocomposite and the graphene nanopowder effectively bridged and sealed shale pore throats, demonstrating superior inhibition performance compared to conventional WBDF. Post adsorption, the shale surface exhibited increased hydrophobicity, contributing to enhanced stability. Overall, the silica nanocomposite and the graphene nanopowder positively impacted rheological performance and provided favorable filtration control in water-based drilling fluids. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of modified polymer-based silica nanocomposite [<a href="#B16-processes-12-01676" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>SEM picture of SiO<sub>2</sub>-NPs (<b>a</b>) and graphene nanopowder (<b>b</b>).</p>
Full article ">Figure 3
<p>OFITE 800 rotational viscosimeter.</p>
Full article ">Figure 4
<p>OFITE HTHP filter press.</p>
Full article ">Figure 5
<p>OFITE dynamic linear swellmeter.</p>
Full article ">Figure 6
<p>FT-IR spectra of the silica nanocomposite.</p>
Full article ">Figure 7
<p>PSD analysis of the diluted silica nanocomposite.</p>
Full article ">Figure 8
<p>TEM image of the diluted silica nanocomposite.</p>
Full article ">Figure 9
<p>FESEM micrograph of WBDF: (<b>a</b>) the base WBDF; (<b>b</b>) the silica nanocomposite WBDF; (<b>c</b>) graphene nanopowder WBDF.</p>
Full article ">
17 pages, 2461 KiB  
Article
New Approaches for Basophil Activation Tests Employing Dendrimeric Antigen–Silica Nanoparticle Composites
by Silvia Calvo-Serrano, Esther Matamoros, Jose Antonio Céspedes, Rubén Fernández-Santamaría, Violeta Gil-Ocaña, Ezequiel Perez-Inestrosa, Cecilia Frecha, Maria I. Montañez, Yolanda Vida, Cristobalina Mayorga and Maria J. Torres
Pharmaceutics 2024, 16(8), 1039; https://doi.org/10.3390/pharmaceutics16081039 - 3 Aug 2024
Viewed by 708
Abstract
In vitro cell activation through specific IgE bound to high-affinity receptors on the basophil surface is a widely used strategy for the evaluation of IgE-mediated immediate hypersensitivity reactions to betalactams. Cellular activation requires drug conjugation to a protein to form a large enough [...] Read more.
In vitro cell activation through specific IgE bound to high-affinity receptors on the basophil surface is a widely used strategy for the evaluation of IgE-mediated immediate hypersensitivity reactions to betalactams. Cellular activation requires drug conjugation to a protein to form a large enough structure displaying a certain distance between haptens to allow the cross-linking of two IgE antibodies bound to the basophil’s surface, triggering their degranulation. However, no information about the size and composition of these conjugates is available. Routine in vitro diagnosis using the basophil activation test uses free amoxicillin, which is assumed to conjugate to a carrier present in blood. To standardize the methodology, we propose the use of well-controlled and defined nanomaterials functionalized with amoxicilloyl. Silica nanoparticles decorated with PAMAM–dendrimer–amoxicilloyl conjugates (NpDeAXO) of different sizes and amoxicilloyl densities (50–300 µmol amoxicilloyl/gram nanoparticle) have been prepared and chemically characterized. Two methods of synthesis were performed to ensure reproducibility and stability. Their functional effect on basophils was measured using an in-house basophil activation test (BAT) that determines CD63+ or CD203chigh activation markers. It was observed that NpDeAXO nanocomposites are not only able to specifically activate basophils but also do so in a more effective way than free amoxicillin, pointing to a translational potential diagnosis. Full article
Show Figures

Figure 1

Figure 1
<p>General procedure for the chemical modification of ϕ<b>dNp</b> dispersions and <b>50Nps</b> with different DeAXO densities in their surface.</p>
Full article ">Figure 2
<p>NMR spectra of (<b>a</b>) AX in basic D<sub>2</sub>O and (<b>b</b>) <b>50NpDeAXO</b> in D<sub>2</sub>O suspensions.</p>
Full article ">Figure 3
<p>Basophil activation test (BAT) dose–response curves of <b>NpDeAXO</b> of different sizes: 20 nm (<b>A</b>,<b>B</b>), 30 nm (<b>C</b>,<b>D</b>), and 50 nm (<b>E</b>,<b>F</b>) and (<b>G</b>,<b>H</b>). <span class="html-italic">Np dispersions</span> or <span class="html-italic">solid-state Nps</span> labels at the top of the figure only indicate the synthetic methodology used for Np preparation. Black lines represent healthy controls (HCs), and blue and green lines represent allergic patients (APs). Size sample included HCs (N = 10) and APs (N = 10) for Nps synthesized as dispersions (<b>A</b>–<b>F</b>), while HCs (N = 45) and APs (N = 54) were included in the study for the Nps synthesized as a solid state. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 4
<p>BAT dose–response curves of Nps with different DeAXO surface densities: 300 µmol AXO/gNp (<b>A</b>,<b>B</b>); 130 µmolAXO/gNp (<b>C</b>,<b>D</b>); 100 µmol AXO/gNp (<b>E</b>,<b>F</b>); 80 µmol AXO/gNp (<b>G</b>,<b>H</b>); 50 µmol AXO/gNp (<b>I</b>,<b>J</b>). Black lines represent healthy controls (HCs), and blue and green lines represent allergic patients (APs). Size sample included HCs (N = 6) and APs (N = 4) (<b>A</b>–<b>J</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01.</p>
Full article ">Figure 5
<p>BAT dose–response curves using <b>50NpDeAXO</b> and free AX at three different concentrations. APs (N = 54) are depicted by blue or green lines and HCs (N = 45) are depicted by black lines. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, *** <span class="html-italic">p</span> &lt; 0.001, **** <span class="html-italic">p</span> &lt; 0.0001.</p>
Full article ">Figure 6
<p>Percentage of positive cases in allergic patients (N = 54) in BAT using CD63 (<b>A</b>) and CD203c<sup>high</sup> (<b>B</b>) as basophil activation markers. Positive cases were obtained after using the cut-offs described in <a href="#app1-pharmaceutics-16-01039" class="html-app">Table S4</a> for AX and <b>50NpDeAXO</b> at the different concentrations with both CD63 and CD203c.</p>
Full article ">
26 pages, 3870 KiB  
Article
Circular Economy of Construction and Demolition Waste for Nanocomposite Cement: XRD, NMR, Vickers, Voltammetric and EIS Characterization
by Roxana Rada, Daniela Lucia Manea, Simona Rada and Radu Fechete
Nanomaterials 2024, 14(15), 1239; https://doi.org/10.3390/nano14151239 - 23 Jul 2024
Viewed by 521
Abstract
In this paper, we present the structural, mechanical and electrical properties of composite cement materials that can be widely used as substituent for cement. We start with the characterization of a composite cement sample using an analysis of X-ray diffraction (XRD) and nuclear [...] Read more.
In this paper, we present the structural, mechanical and electrical properties of composite cement materials that can be widely used as substituent for cement. We start with the characterization of a composite cement sample using an analysis of X-ray diffraction (XRD) and nuclear magnetic resonance (NMR) spectra. The measurements of the Vickers hardness, cyclic and sweep linear voltammetry and electrochemical impedance spectroscopy (EIS) of composite cement materials were also recorded. This study compared the effect of the different nanocomposites added to cement on the mitigation of the alkali–silica reaction, which is responsible for the swelling, cracking and deleterious behavior of the material. The enhancement in Vickers hardness was more pronounced for composite cement materials. In contrast, the values of Vickers hardness decreased for the composite cement containing mortar and the control sample, suggesting that the long-term performance of cement was compromised. In order to obtain information about the bulk resistance of the composite cement material, electrochemical impedance spectroscopy (EIS) data were employed. The results suggest that for composite cement materials, there is an improvement in bulk electrical resistance, which can be attributed to the lower amounts of cracks and swelling due to lower expansion. In the control sample, a reduction in the bulk resistance suggests the formation of microcracks, which cause the aging and degradation of the material. The intersection of arcs in the EIS spectrum of the mixed composite cement sample gradually increased by an alkaline exposure of up to 21 days and finally shifted towards a low value of high frequency with an increase in alkaline exposure of up to 28 days. Full article
Show Figures

Figure 1

Figure 1
<p>X-ray patterns of the composite—cement materials in the region between (<b>a</b>) 10–60 degrees and (<b>b</b>) 28–35 degrees. The Miller indices are also inserted in the subfigure.</p>
Full article ">Figure 2
<p>Compositional evolution of average particles sizes from composite–cement.</p>
Full article ">Figure 3
<p>NMR spectra of the composite–cement samples recorded at (<b>a</b>) 3, (<b>b</b>) 7, (<b>c</b>) 14 and (<b>d</b>) 28 days after their preparation. The types of water reservoirs are also inserted.</p>
Full article ">Figure 3 Cont.
<p>NMR spectra of the composite–cement samples recorded at (<b>a</b>) 3, (<b>b</b>) 7, (<b>c</b>) 14 and (<b>d</b>) 28 days after their preparation. The types of water reservoirs are also inserted.</p>
Full article ">Figure 3 Cont.
<p>NMR spectra of the composite–cement samples recorded at (<b>a</b>) 3, (<b>b</b>) 7, (<b>c</b>) 14 and (<b>d</b>) 28 days after their preparation. The types of water reservoirs are also inserted.</p>
Full article ">Figure 3 Cont.
<p>NMR spectra of the composite–cement samples recorded at (<b>a</b>) 3, (<b>b</b>) 7, (<b>c</b>) 14 and (<b>d</b>) 28 days after their preparation. The types of water reservoirs are also inserted.</p>
Full article ">Figure 4
<p>The relaxation time distributions of mobile water from NMR data of composite–cement materials at. (<b>a</b>) interlayer peak, (<b>b</b>) gel pore peak, (<b>c</b>) inter-hydrate pore peak, and (<b>d</b>) capillarity pore peak at 28 days after their preparation.</p>
Full article ">Figure 5
<p>(<b>a</b>) Microscopic images after indentation (in the red square is shown the pyramidal area of the indentation) and (<b>b</b>) compositional dependence of Vickers hardness distributions of composite–cement materials (with red dashed line indicating the Vickers value of the control sample).</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Microscopic images after indentation (in the red square is shown the pyramidal area of the indentation) and (<b>b</b>) compositional dependence of Vickers hardness distributions of composite–cement materials (with red dashed line indicating the Vickers value of the control sample).</p>
Full article ">Figure 6
<p>EIS spectrum for (<b>a</b>) perfect and (<b>b</b>) depressed semicircle. The equivalent circuits are also shown.</p>
Full article ">Figure 7
<p>(<b>a</b>) Nyquist curves in an alkaline solution and (<b>b</b>) the bulk electric resistance, R<sub>b</sub>, as a function of composite–cement composition. The red line shows the value of the control specimen.</p>
Full article ">Figure 8
<p>Cyclic voltammograms recorded in alkaline solution of composite–cement materials (control and all samples) using working electrode for one cycle (<b>a</b>,<b>b</b>) and after the scanning of the three cycles (<b>c</b>).</p>
Full article ">Figure 8 Cont.
<p>Cyclic voltammograms recorded in alkaline solution of composite–cement materials (control and all samples) using working electrode for one cycle (<b>a</b>,<b>b</b>) and after the scanning of the three cycles (<b>c</b>).</p>
Full article ">Figure 9
<p>(<b>a</b>) Sweep linear voltammograms and (<b>b</b>) the values of the half-wave potentials, E<sub>1/2</sub>, versus composite–cement composition (the red line indicates the value of the control sample).</p>
Full article ">Figure 10
<p>Nyquist plots of (<b>a</b>) control cement and (<b>b</b>) all composites—cement sample. (<b>c</b>) Resistance values recorded after some days of alkaline environment of the studied cementitious materials.</p>
Full article ">Figure 10 Cont.
<p>Nyquist plots of (<b>a</b>) control cement and (<b>b</b>) all composites—cement sample. (<b>c</b>) Resistance values recorded after some days of alkaline environment of the studied cementitious materials.</p>
Full article ">Figure 10 Cont.
<p>Nyquist plots of (<b>a</b>) control cement and (<b>b</b>) all composites—cement sample. (<b>c</b>) Resistance values recorded after some days of alkaline environment of the studied cementitious materials.</p>
Full article ">Figure 11
<p>(<b>a</b>) The evolution of scientific publications on construction and demolition waste, their recycling and circular economy. (<b>b</b>) The main years with the highest percentage of publications on the circular economy of C&amp;D waste.</p>
Full article ">
13 pages, 4243 KiB  
Article
An Insight into the Varying Effects of Different Cryogenic Temperatures on the Microstructure and the Thermal and Compressive Response of a Mg/SiO2 Nanocomposite
by Michael Johanes, Sarah Mehtabuddin, Vishal Venkatarangan and Manoj Gupta
Metals 2024, 14(7), 808; https://doi.org/10.3390/met14070808 - 11 Jul 2024
Viewed by 529
Abstract
This study for the first time reports that insights into microstructure and thermal and compressive responses can be best achieved following exposure to different cryogenic temperatures and that the lowest cryogenic temperature may not always produce the best results. In the present study, [...] Read more.
This study for the first time reports that insights into microstructure and thermal and compressive responses can be best achieved following exposure to different cryogenic temperatures and that the lowest cryogenic temperature may not always produce the best results. In the present study, a Mg-SiO2 biocompatible and environment-friendly nanocomposite was synthesized by using the Disintegrated Melt Deposition method followed by hot extrusion. Subsequently, it was subjected to four different sub-zero temperatures (−20 °C, −50 °C, −80 °C, and −196 °C). The results reveal the best densification at −80 °C, marginally improved ignition resistance at 50 °C, the best damping response at −80 °C, the best microhardness at −50 °C, and the best compressive response at −20 °C. The results clearly indicate that the cryogenic temperature should be carefully chosen depending on the property that needs to be particularly enhanced governed by the principal requirement of the end application. Full article
(This article belongs to the Special Issue Design and Development of Metal Matrix Composites)
Show Figures

Figure 1

Figure 1
<p>Scanning electron micrographs of Mg-2SiO<sub>2</sub> materials in this work, with selected regions used for the EDS studies.</p>
Full article ">Figure 2
<p>EDS results/chart of Mg-2SiO<sub>2</sub> AE: matrix region.</p>
Full article ">Figure 3
<p>X-ray diffractograms of Mg-2SiO<sub>2</sub> materials in this work.</p>
Full article ">Figure 4
<p>The ignition responses of Mg-2SiO<sub>2</sub> materials in this work, showing the locations where ignition temperatures were evaluated.</p>
Full article ">Figure 5
<p>Thermal response of Mg-2SiO<sub>2</sub> materials in this work.</p>
Full article ">Figure 6
<p>Macro-scale photographs of compressed Mg-2SiO<sub>2</sub> materials in this work.</p>
Full article ">Figure 7
<p>Fractographs of Mg-2SiO<sub>2</sub> materials in this work.</p>
Full article ">
19 pages, 5723 KiB  
Article
Synthesis of TiO2/SBA-15 Nanocomposites by Hydrolysis of Organometallic Ti Precursors for Photocatalytic NO Abatement
by Ons El Atti, Julie Hot, Katia Fajerwerg, Christian Lorber, Bénédicte Lebeau, Andrey Ryzhikov, Myrtil Kahn, Vincent Collière, Yannick Coppel, Nicolas Ratel-Ramond, Philippe Ménini and Pierre Fau
Inorganics 2024, 12(7), 183; https://doi.org/10.3390/inorganics12070183 - 29 Jun 2024
Viewed by 648
Abstract
The development of advanced photocatalysts for air pollution removal is essential to improve indoor air quality. TiO2/mesoporous silica SBA-15 nanocomposites were synthesized using an organometallic decoration method, which leverages the high reactivity of Ti precursors to be hydrolyzed on the surface [...] Read more.
The development of advanced photocatalysts for air pollution removal is essential to improve indoor air quality. TiO2/mesoporous silica SBA-15 nanocomposites were synthesized using an organometallic decoration method, which leverages the high reactivity of Ti precursors to be hydrolyzed on the surface water groups of silica supports. Both lab-made Ti(III) amidinate and commercial Ti(IV) amino precursors were utilized to react with water-rich SBA-15, obtained through a hydration process. The hydrated SBA-15 and the TiO2/SBA-15 nanocomposites were characterized using TGA, FTIR, 1H and 29Si NMR, TEM, SEM, N2 physisorption, XRD, and WAXS. This one-step TiO2 decoration method achieved a loading of up to 51.5 wt.% of approximately 9 nm anatase particles on the SBA-15 surface. This structuring provided excellent accessibility of TiO2 particles for photocatalytic applications under pollutant gas and UV-A light exposure. The combination with the high specific surface area of SBA-15 resulted in the efficient degradation of 400 ppb of NO pollutant gas. Due to synergistic effects, the best nanocomposite in this study demonstrated a NO abatement performance of 4.0% per used mg of TiO2, which is 40% more efficient than the reference photocatalytic material TiO2 P-25. Full article
(This article belongs to the Special Issue Feature Papers in Inorganic Materials 2024)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>TGA analysis of SBA-15 powders according to hydration methods. (<b>a</b>) As-received; (<b>b</b>) exposed to air atmosphere with 75% RH for 4 h (method A); and (<b>c</b>) soaked in boiling water for 2 h (method B).</p>
Full article ">Figure 2
<p><sup>1</sup>H NMR MAS spectra of SBA-15 powders according to hydration methods. (<b>a</b>) As-received; (<b>b</b>) exposed to air with 75% RH for 4 h (method A); and (<b>c</b>) soaked in boiling water for 2 h (method B).</p>
Full article ">Figure 3
<p>Comparison of PDF obtained from WAXS measurement of TiOx powders from TEMAT (<b>1</b>, red dotted line) and Ti-Amd (<b>2</b>, plain black line) precursors according to the calcination temperature at: (<b>a</b>) 150 °C, (<b>b</b>) 350 °C, and (<b>c</b>) 500 °C. Refinement of the pair distribution function obtained on TiO<sub>2</sub>-amd calcinated at 350 °C with (<b>d</b>) brookite and (<b>e</b>) anatase structures.</p>
Full article ">Figure 4
<p>TEM images (left) and corresponding histograms of particle size distribution (right) for TiO<sub>2</sub> powders (determined by ImageJ software (v1.54j) on around 100 particles) after calcination at 500 °C from the hydrolysis of (<b>a</b>) TEMAT (<b>1</b>) and (<b>b</b>) Ti-Amd (<b>2</b>).</p>
Full article ">Figure 5
<p>SEM images of TiO<sub>2</sub>/SBA-15 calcined at 500 °C and prepared with (<b>a</b>) TEMAT (<b>1</b>) and (<b>b</b>) Ti-Amd (<b>2</b>). The images are in chemical contrast mode (back-scattered electrons). The brighter spots on the image correspond to an element with a higher atomic number and indicate the presence of Ti. Magnification is ×5000.</p>
Full article ">Figure 6
<p>SEM images of TiO<sub>2</sub>/SBA-15 prepared with (<b>a</b>) TEMAT (<b>1</b>) (top) and images in artificial color corresponding to the EDS chemical mapping of elements Ti, Si, and O from the sample (bottom); and (<b>b</b>) Ti-Amd (<b>2</b>) (top) and EDS chemical mapping for Ti, Si, and O (bottom). Ti atoms are shown in pink, Si atoms are shown in blue, and O atoms are shown in green. A flash platinum layer is deposited on the sample in order to enhance its electrical conductivity and improve the image quality (Pt, red dots displayed uniformly on the image). Magnification is ×5000 for (<b>a</b>) and ×10,000 for (<b>b</b>).</p>
Full article ">Figure 7
<p>TEM images of TiO<sub>2</sub>/SBA-15 nanocomposites obtained from hydrolysis of TEMAT (<b>1</b>) (<b>a</b>) magnification ×10,000, (<b>b</b>) magnification ×100,000, and from hydrolysis of Ti-Amd (<b>2</b>) (<b>c</b>) magnification ×10,000, (<b>d</b>) magnification ×100,000.</p>
Full article ">Figure 8
<p>Schematic description of the reaction hydrolysis of Ti-Amd precursor on hydrated SBA-15.</p>
Full article ">Figure 9
<p>Degradation of NO under UV-A artificial light for the same surface density of TiO<sub>2</sub> (0.02 mg of TiO<sub>2</sub> per cm<sup>2</sup> of glass surface) obtained with bare TiO<sub>2</sub> (TiO<sub>2</sub> synthesized from <b>1</b> (TEMAT) and <b>2</b> (Ti-Amd) calcined at 500 °C, and TiO<sub>2</sub> P-25), TiO<sub>2</sub>/SBA-15 nanocomposites (calcined at 500 °C) and SBA-15 physically mixed with bare TiO<sub>2</sub> oxides (calcined at 500 °C and obtained from <b>1</b> and <b>2</b>).</p>
Full article ">Scheme 1
<p>Hydrolysis reaction of precursor <b>1</b> or <b>2</b> with a stoichiometric amount of water.</p>
Full article ">
29 pages, 3055 KiB  
Review
Liquid Nanoclay: Synthesis and Applications to Transform an Arid Desert into Fertile Land
by Kamel A. Abd-Elsalam, Mirza Abid Mehmood, Muhammad Ashfaq, Toka E. Abdelkhalek, Rawan K. Hassan and Mythili Ravichandran
Soil Syst. 2024, 8(3), 73; https://doi.org/10.3390/soilsystems8030073 - 27 Jun 2024
Viewed by 1358
Abstract
Nanoclay, a processed clay, is utilized in numerous high-performance cement nanocomposites. This clay consists of minerals such as kaolinite, illite, chlorite, and smectite, which are the primary components of raw clay materials formed in the presence of water. In addition to silica, alumina, [...] Read more.
Nanoclay, a processed clay, is utilized in numerous high-performance cement nanocomposites. This clay consists of minerals such as kaolinite, illite, chlorite, and smectite, which are the primary components of raw clay materials formed in the presence of water. In addition to silica, alumina, and water, it also contains various concentrations of inorganic ions like Mg2+, Na+, and Ca2+. These are categorized as hydrous phyllosilicates and can be located either in interlayer spaces or on the planetary surface. Clay minerals are distinguished by their two-dimensional sheets and tetrahedral (SiO4) and octahedral (Al2O3) crystal structures. Different clay minerals are classified based on the presence of tetrahedral and octahedral layers in their structure. These include kaolinite, which has a 1:1 ratio of tetrahedral to octahedral layers, the smectite group of clay minerals and chlorite with a 2:1 ratio. Clay minerals are unique due to their small size, distinct crystal structure, and properties such as high cation exchange capacity, adsorption capacity, specific surface area, and swelling behavior. These characteristics are discussed in this review. The use of nanoclays as nanocarriers for fertilizers boasts a diverse array of materials available in both anionic and cationic variations. Layered double hydroxides (LDH) possess a distinctive capacity for exchanging anions, making them suitable for facilitating the transport of borate, phosphate, and nitrate ions. Liquid nanoclays are used extensively in agriculture, specifically as fertilizers, insecticides, herbicides, and nutrients. These novel nanomaterials have numerous benefits, including improved nutrient use, controlled nutrient release, targeted nutrient delivery, and increased agricultural productivity. Arid regions face distinct challenges like limited water availability, poor soil quality, and reduced productivity. The addition of liquid nanoclay to sandy soil offers a range of benefits that contribute to improved soil quality and environmental sustainability. Liquid nanoclay is being proposed for water management in arid regions, which will necessitate a detailed examination of soil, water availability, and hydrological conditions. Small-scale trial initiatives, engagement with local governments, and regular monitoring are required to fully comprehend its benefits and drawbacks. These developments would increase the practicality and effectiveness of using liquid nanoclay in desert agriculture. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Chemical structures of 1:1 (<b>A</b>) and 2:1 (<b>B</b>) phyllosilicates [<a href="#B61-soilsystems-08-00073" class="html-bibr">61</a>].</p>
Full article ">Figure 2
<p>The 2:1:1 layer phyllosilicate structure [<a href="#B61-soilsystems-08-00073" class="html-bibr">61</a>].</p>
Full article ">Figure 3
<p>Distinctive properties of nanoclay-based pickering emulsions for applications in various industries.</p>
Full article ">Figure 4
<p>Liquid nanoclay is produced by dispersing a layered <span class="html-italic">clay</span> in water through an innovative mixing method that generates laminar and turbulent flow regimes. This dispersion occurs because the cationic nature of the nanoclay particles attracts water molecules, which act as anions, surrounding each flake. (<b>A</b>) Powdered clay. (<b>B</b>) Clay dissolved in water. (<b>C</b>) <span class="html-italic">Hylocereus undatus</span> cultivated in sand soil irrigated with LNC [<a href="#B71-soilsystems-08-00073" class="html-bibr">71</a>].</p>
Full article ">Figure 5
<p>The field with (<b>right</b>) and without (<b>left</b>) liquid nanoclay (LNC) treatment illustrated LNC’s impact on soil, nutrient, water management, and environmental sustainability.</p>
Full article ">Figure 6
<p>A schematic overview of nanoclay vehiculization for eco-friendly and economically viable agriculture techniques.</p>
Full article ">Figure 7
<p>Estimated capital cost (<b>a</b>) and annual operational cost (<b>b</b>) for nanoclay from bentonite.</p>
Full article ">Figure 8
<p>Liquid nanoclay (LNC) formulation and its long-term effectiveness potential under field conditions.</p>
Full article ">
21 pages, 8420 KiB  
Review
In Situ Generation of Nanoparticles on and within Polymeric Materials
by Antonios Kelarakis
Polymers 2024, 16(11), 1611; https://doi.org/10.3390/polym16111611 - 6 Jun 2024
Viewed by 865
Abstract
It is well-established that the structural, morphological and performance characteristics of nanoscale materials critically depend upon the dispersion state of the nanofillers that is, in turn, largely determined by the preparation protocol. In this report, we review synthetic strategies that capitalise on the [...] Read more.
It is well-established that the structural, morphological and performance characteristics of nanoscale materials critically depend upon the dispersion state of the nanofillers that is, in turn, largely determined by the preparation protocol. In this report, we review synthetic strategies that capitalise on the in situ generation of nanoparticles on and within polymeric materials, an approach that relies on the chemical transformation of suitable precursors to functional nanoparticles synchronous with the build-up of the nanohybrid systems. This approach is distinctively different compared to standard preparation methods that exploit the dispersion of preformed nanoparticles within the macromolecular host and presents advantages in terms of time and cost effectiveness, environmental friendliness and the uniformity of the resulting composites. Notably, the in situ-generated nanoparticles tend to nucleate and grow on the active sites of the macromolecular chains, showing strong adhesion on the polymeric host. So far, this strategy has been explored in fabrics and membranes comprising metallic nanoparticles (silver, gold, platinum, copper, etc.) in relation to their antimicrobial and antifouling applications, while proof-of-concept demonstrations for carbon- and silica-based nanoparticles as well as titanium oxide-, layered double hydroxide-, hectorite-, lignin- and hydroxyapatite-based nanocomposites have been reported. The nanocomposites thus prepared are ideal candidates for a broad spectrum of applications such as water purification, environmental remediation, antimicrobial treatment, mechanical reinforcement, optical devices, etc. Full article
(This article belongs to the Section Smart and Functional Polymers)
Show Figures

Figure 1

Figure 1
<p>Scanning Electron Microscopy (SEM) images of (<b>a</b>) the surface and (<b>b</b>) the interior of bleached cotton fibres following the <span class="html-italic">in situ</span> formation of Ag NPs. (<b>c</b>) Outlet concentration of Ag as a function of time when <span class="html-italic">in situ</span>-modified wool fibres were subjected to flow-through experiments using the inlet solutions indicated (flow rate 1 mL/min). Adapted with permission from Ref. [<a href="#B29-polymers-16-01611" class="html-bibr">29</a>].</p>
Full article ">Figure 2
<p>Silk fabric bearing <span class="html-italic">in situ</span>-generated Ag NPs: (<b>a</b>) their colour with respect to the AgNO<sub>3</sub> and FA concentration; (<b>b</b>) their antibacterial activity against <span class="html-italic">E. coli</span> when subjected to a number of washing cycles. Adapted with permission from Ref. [<a href="#B33-polymers-16-01611" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Porous cellulose fibres bearing <span class="html-italic">in situ</span>-generated Ag, Au, Pt, and Pd NPs compared to the neat fibres. Transmission Electron Microscopy (TEM) images of porous cellulose fibres bearing <span class="html-italic">in situ</span>-generated Ag (<b>b</b>), Au (<b>c</b>), Pt (<b>d</b>), and Pd (<b>e</b>) NPs prepared from aqueous dispersions AgNO<sub>3</sub>, AuCl<sub>3</sub>, PtCl<sub>4</sub> and Pd(NO<sub>3</sub>)<sub>2</sub>, respectively. Adapted with permission from Ref. [<a href="#B37-polymers-16-01611" class="html-bibr">37</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Percentage of Ag NPs remaining immobilised on the surface of TA-Fe-PEI/Ag-modified membrane following its immersion in aqueous NaHCO<sub>3</sub> with pH = 8.2. Photos of the Petri dishes containing <span class="html-italic">B. subtilis</span> (<b>b</b>,<b>c</b>) and <span class="html-italic">E. coli</span> (<b>d</b>,<b>e</b>) cultures exposed for 1.5 h to unmodified PA (<b>c</b>,<b>e</b>) and TA-Fe-PEI/Ag-modified membrane (<b>b</b>,<b>d</b>). Adapted with permission from Ref. [<a href="#B39-polymers-16-01611" class="html-bibr">39</a>].</p>
Full article ">Figure 5
<p><span class="html-italic">In situ</span>-modified PA/Ag NPs membranes: (<b>a</b>,<b>b</b>) TEM images; reduction in the bacterial population of <span class="html-italic">E. coli</span> (<b>c</b>), <span class="html-italic">P. aeruginosa</span> (<b>d</b>) and <span class="html-italic">S. aureus</span> (<b>e</b>) exposed to <span class="html-italic">in situ</span>-modified membranes (green) compared to their pristine counterparts (grey). (<b>f</b>) Biovolumes of live, dead and extracellular polymeric substances (EPS) images of <span class="html-italic">P. aeruginosa</span> biofilm grown for 24 h on pristine and <span class="html-italic">in situ</span>-modified PA membranes. Asterisks represent significant (<span class="html-italic">p</span> &lt; 0.05) difference between groups. Adapted with permission from Ref. [<a href="#B40-polymers-16-01611" class="html-bibr">40</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Water flux and NaCl rejection of the <span class="html-italic">in situ</span>-modified PES/PA-TA-Ag membranes compared to their unmodified counterparts. (<b>b</b>) Average biofilm thickness deposited on PES/PA-TA-Ag membranes. Adapted with permission from Ref. [<a href="#B41-polymers-16-01611" class="html-bibr">41</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) PNP percental conversion in a flow membrane reactor with PET/DOPA/PEI/Au(3 h), PET/DOPA/PEI/Au(6 h) and PET/DOPA/PEI/Au(9 h) membranes as a function of flux. The inset displays data for PET/DOPA/Au(6 h). (<b>b</b>) Absorbance decrease at 495 nm during the degradation of Congo red under static conditions in the presence of PET/DOPA/PEI/Au(9 h) membrane using NaBH<sub>4</sub> as the reducing agent as a function of time. (<b>c</b>) Degradation percentage of Congo red in a flow reactor with PET/DOPA/PEI/Au(9 h) over 11 cycles. Adapted with permission from Ref. [<a href="#B44-polymers-16-01611" class="html-bibr">44</a>].</p>
Full article ">Figure 8
<p>PVA/<span class="html-italic">in situ</span> ZnO:CS electrospun fibres (PVA/ZnO:CS ratio 40/60): (<b>a</b>) SEM image, (<b>b</b>) TEM image and (<b>c</b>) nanoindentation tests. In all cases, the applied voltage was 25 kV, the needle-tip collector distance was 20 cm, and the flow rate was 3 μL/s. Adapted with permission from Ref. [<a href="#B53-polymers-16-01611" class="html-bibr">53</a>].</p>
Full article ">Figure 9
<p>(<b>a</b>) Radical scavenging activity and (<b>b</b>) photocatalytic degradation activity against MB of the nanocomposites fibrous membranes as a function of time. Adapted with permission from Ref. [<a href="#B54-polymers-16-01611" class="html-bibr">54</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) TEM images (showing particle size histogram and lattice plane distance in Ag NPs) of the PA6-based nanocomposite prepared with 20 wt% Ag<sub>2</sub>O that was melt processed for 10 min at 240 °C. (<b>b</b>) Contact killing activity against <span class="html-italic">L. monocytogenes</span> of PA6 films containing 0.5 wt % of AgNP compared to the neat polymer (control). (<b>c</b>) Agar diffusion tests using PA6 nanocomposites with various Ag NPs loadings. Adapted with permission from Ref. [<a href="#B55-polymers-16-01611" class="html-bibr">55</a>].</p>
Full article ">Figure 11
<p>SEM images for CNT grown <span class="html-italic">in situ</span> on (<b>a</b>) Teflon and (<b>b</b>) polycarbonate. Adapted with permission from Ref. [<a href="#B59-polymers-16-01611" class="html-bibr">59</a>].</p>
Full article ">Figure 12
<p>(<b>a</b>) TEM image of the C-dots extracted from PEG/C-dot nanocomposite. (<b>b</b>) <sup>1</sup>H NMR spectra of PEG/C-dot, PEG and EA in D<sub>2</sub>O. Fluorescence microscopy images of (<b>c</b>) <span class="html-italic">in situ</span> generated PP/C-dot and (<b>d</b>) PEG/C-dot nanocomposite under (i) UV violet, (ii) blue and (iii) green excitation wavelength. Adapted with permission from Ref. [<a href="#B64-polymers-16-01611" class="html-bibr">64</a>].</p>
Full article ">Figure 13
<p>SEM images of: (<b>a</b>–<b>d</b>) POSS layer deposited of Kapton film at different growth times (2 h, 4 h, 6 h and 8 h), (<b>e</b>–<b>h</b>) sputter-deposited SiO<sub>2</sub> coatings on different POSS/Kapton films. Scale bars represent 200 nm. Adapted with permission from Ref. [<a href="#B70-polymers-16-01611" class="html-bibr">70</a>].</p>
Full article ">Figure 14
<p>(<b>a</b>) SEM image of a freeze-cast Nafion/GO/H<sub>2</sub>PtCl<sub>6</sub> microporous scaffold showing the presence of GO nanosheets (arrow) on the surface of the macropores. (<b>b</b>) Pt NPs deposited on the surface of graphene, after hydrazine treatment, and (<b>c</b>) X-ray diffraction (XRD) pattern of Pt NPs. (<b>d</b>,<b>e</b>) SEM and TEM images, respectively, of Pt NPs deposited on the surface of graphene, after treatment with sodium citrate. Adapted with permission from Ref. [<a href="#B78-polymers-16-01611" class="html-bibr">78</a>].</p>
Full article ">Figure 15
<p>(<b>a</b>) TEM image, (<b>b</b>) size distribution and (<b>c</b>) HRTEM of <span class="html-italic">in situ</span>-generated PAN/Ag nanocomposite. Adapted with permission from Ref. [<a href="#B79-polymers-16-01611" class="html-bibr">79</a>].</p>
Full article ">Figure 16
<p>PPy/Ag/clay nanocomposites (mass ratio of ATP to pyrrole was 20:100): (<b>a</b>) SEM, (<b>b</b>) TEM images and (<b>c</b>) TGA. Adapted with permission from Ref. [<a href="#B80-polymers-16-01611" class="html-bibr">80</a>].</p>
Full article ">Figure 17
<p>(<b>a</b>) TEM images of C-dots extracted from the PDMS/C-dot1 nanocomposite. The scale bars correspond to 10 nm (<b>left</b> image) and 2 nm (<b>right</b> image). (<b>b</b>) Photos of PDMS/C-dot1, PDMS/Cdot2 and PDMS/C-dot3 nanocomposites under illumination at 365 nm. The numbers refer to the C-dot precursors 1,2,3 corresponding to 6-O-(O-O′-Di-lauroyl-tartaryl)-d-glucose, 6-O-(O-O′-Di-lauroyl-tartaryl)-l-ascorbic acid and Vitamin B1 + oleic acid, respectively. Adapted with permission from Ref. [<a href="#B82-polymers-16-01611" class="html-bibr">82</a>].</p>
Full article ">
21 pages, 3397 KiB  
Article
In Vitro Assessment of 177Lu-Labeled Trastuzumab-Targeted Mesoporous Carbon@Silica Nanostructure for the Treatment of HER2-Positive Breast Cancer
by Ayça Tunçel, Simone Maschauer, Olaf Prante and Fatma Yurt
Pharmaceuticals 2024, 17(6), 732; https://doi.org/10.3390/ph17060732 - 5 Jun 2024
Viewed by 775
Abstract
This study assessed the effectiveness of a trastuzumab-targeted 177Lu-labeled mesoporous Carbon@Silica nanostructure (DOTA@TRA/MC@Si) for HER2-positive breast cancer treatment, focusing on its uptake, internalization, and efflux in breast cancer cells. The synthesized PEI-MC@Si nanocomposite was reacted with DOTA-NHS-ester, confirmed by the Arsenazo(III) assay. [...] Read more.
This study assessed the effectiveness of a trastuzumab-targeted 177Lu-labeled mesoporous Carbon@Silica nanostructure (DOTA@TRA/MC@Si) for HER2-positive breast cancer treatment, focusing on its uptake, internalization, and efflux in breast cancer cells. The synthesized PEI-MC@Si nanocomposite was reacted with DOTA-NHS-ester, confirmed by the Arsenazo(III) assay. Following this, TRA was conjugated to the DOTA@PEI-MC@Si for targeting. DOTA@PEI-MC@Si and DOTA@TRA/MC@Si nanocomposites were labeled with 177Lu, and their efficacy was evaluated through in vitro radiolabeling experiments. According to the results, the DOTA@TRA/MC@Si nanocomposite was successfully labeled with 177Lu, yielding a radiochemical yield of 93.0 ± 2.4%. In vitro studies revealed a higher uptake of the [177Lu]Lu-DOTA@TRA/MC@Si nanocomposite in HER2-positive SK-BR-3 cells (44.0 ± 4.6% after 24 h) compared to MDA-MB-231 cells (21.0 ± 2.3%). The IC50 values for TRA-dependent uptake in the SK-BR-3 and BT-474 cells were 0.9 µM and 1.3 µM, respectively, indicating affinity toward HER-2 receptor-expressing cells. The lipophilic distribution coefficients of the radiolabeled nanocomposites were determined to be 1.7 ± 0.3 for [177Lu]Lu-DOTA@TRA/MC@Si and 1.5 ± 0.2 for [177Lu]Lu-DOTA@PEI-MC@Si, suggesting sufficient passive transport through the cell membrane and increased accumulation in target tissues. The [177Lu]Lu-DOTA@TRA/MC@Si nanocomposite showed an uptake into HER2-positive cell lines, marking a valuable step toward the development of a nanoparticle-based therapeutic agent for an improved treatment strategy for HER2-positive breast cancer. Full article
(This article belongs to the Special Issue Radiopharmaceuticals and Nanotechnology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Absorbance/concentration curve of octylamine drawn with the ninhydrin test. The green dots show different concentrations of Octylamine. (<b>b</b>) Standard calibration graph dependent on the absorption detected at 595 nm for the DOTA-NHS-ester concentration. The red dots indicate various concentrations of DOTA-NHS ester standards. (<b>c</b>) BSA standard absorbance/concentration graph. The blue dots represent different concentrations of BSA standards. Data are presented as mean values ± SD, determined in triplicate from two independent experiments.</p>
Full article ">Figure 2
<p>ITLC chromatograms obtained for [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si in a 10 mM DTPA chamber solution.</p>
Full article ">Figure 3
<p>ITLC chromatograms obtained for [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si in a 0.1 M sodium citrate (pH 5)/water (1:5) chamber solution.</p>
Full article ">Figure 4
<p>ITLC chromatograms obtained for [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si after centrifugation in a 10 mM DTPA chamber solution.</p>
Full article ">Figure 5
<p>Evaluation of reaction parameters for the radiolabeling of DOTA@TRA/MC@Si nanocomposite with <sup>177</sup>Lu and determination of the RCY of the [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si nanocomposite: (<b>a</b>) amount of nanocomposite, (<b>b</b>) incubation time, (<b>c</b>) temperature, (<b>d</b>) pH, and (<b>e</b>) purification by ultracentrifugation. Data are presented as the mean values ± SD, determined in triplicate from four independent experiments.</p>
Full article ">Figure 6
<p>Cellular uptake of [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si and [<sup>177</sup>Lu]Lu-DOTA@PEI-MC@Si nanocomposites in HER2-positive cells (SK-BR-3 &amp; BT-474) and HER2-deficient cells (MDA-MB-231). Data are presented as mean values ± SD, determined in triplicate from two independent experiments.</p>
Full article ">Figure 7
<p>Competitive binding experiments were conducted with [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si and increasing concentrations of trastuzumab using SK-BR-3 and BT-474. Data are expressed as mean values ± SD (<span class="html-italic">n</span> = 4) from two independent experiments.</p>
Full article ">Figure 8
<p>CPM (Radioactivity)/Time Ratios of [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si Nanocomposite in SK-BR-3, BT-474, and MDA-MB-231, and Cell Lines. Red curve: “internalized” is defined as the radioactivity of internalized and cell surface-bound fraction; Blue curve: “acid wash” is presented as the radio-activity of the cell surface-bound fraction”. Data are presented as mean values ± SD, determined in triplicate from 2 independent experiments.</p>
Full article ">Figure 9
<p>Cellular Internalization and Efflux of [<sup>177</sup>Lu]Lu-DOTA@TRA/MC@Si and [<sup>177</sup>Lu]Lu-DOTA@PEI/MC@Si Nanomaterials in SK-BR-3 Cells. Data are presented as mean values ± SD, determined in triplicate from 2 independent experiments.</p>
Full article ">
13 pages, 2862 KiB  
Article
Design of Environmental-Friendly Carbon-Based Catalysts for Efficient Advanced Oxidation Processes
by Xinru Xu, Guochen Kuang, Xiao Jiang, Shuoming Wei, Haiyuan Wang and Zhen Zhang
Materials 2024, 17(11), 2750; https://doi.org/10.3390/ma17112750 - 5 Jun 2024
Viewed by 536
Abstract
Advanced oxidation processes (AOPs) represent one of the most promising strategies to generate highly reactive species to deal with organic dye-contaminated water. However, developing green and cost-effective catalysts is still a long-term goal for the wide practical application of AOPs. Herein, we demonstrated [...] Read more.
Advanced oxidation processes (AOPs) represent one of the most promising strategies to generate highly reactive species to deal with organic dye-contaminated water. However, developing green and cost-effective catalysts is still a long-term goal for the wide practical application of AOPs. Herein, we demonstrated doping cobalt in porous carbon to efficiently catalyze the oxidation of the typically persistent organic pollutant rhodamine B, via multiple reactive species through the activation of peroxymonosulfate (PMS). The catalysts were prepared by facile pyrolysis of nanocomposites with a core of cobalt-loaded silica and a shell of phenolic resin (Co-C/SiO2). It showed that the produced 1O2 could effectively attack the electron-rich functional groups in rhodamine B, promoting its molecular chain breakage and accelerating its oxidative degradation reaction with reactive oxygen-containing radicals. The optimized Co-C/SiO2 catalyst exhibits impressive catalytic performance, with a degradation rate of rhodamine B up to 96.7% in 14 min and a reaction rate constant (k) as high as 0.2271 min−1, which suggested promising potential for its practical application. Full article
(This article belongs to the Special Issue Advanced Catalysts for Energy and Environmental Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the preparation process of the Co-C/SiO<sub>2</sub> catalyst with core-shell structure; (<b>b</b>) SEM and (<b>c</b>) TEM of SiO<sub>2</sub>; (<b>d</b>) SEM and (<b>e</b>) TEM of Co/SiO<sub>2</sub>; (<b>f</b>) SEM and (<b>g</b>) TEM of Co-C/SiO<sub>2</sub>; the average size distribution diagram of (<b>h</b>) SiO<sub>2</sub>; (<b>i</b>) Co/SiO<sub>2</sub> and (<b>j</b>) Co-C/SiO<sub>2</sub>; (<b>k</b>) EDS elemental mapping images of C, N, O, Si, Co in Co-C/SiO<sub>2</sub>.</p>
Full article ">Figure 2
<p>Effects of various factors on RhB removal in the Co-C/SiO<sub>2</sub>-PMS system. (<b>a</b>) Effect of PMS concentration; (<b>b</b>) the corresponding kinetics and (<b>c</b>) the corresponding degradation rate constants (<span class="html-italic">k</span>) of RhB; (<b>d</b>) effect of inorganic anions; (<b>e</b>) the corresponding kinetics and (<b>f</b>) the corresponding degradation rate constants (<span class="html-italic">k</span>) of RhB. Conditions: [Catalyst] = 0.2 g/L, [PMS] = 0.2 g/L, [RhB] = 25 mg/L, pH = 6.6, T = 25 °C.</p>
Full article ">Figure 3
<p>(<b>a</b>) RhB removal in different systems; (<b>b</b>) the corresponding kinetics and (<b>c</b>) the corresponding degradation rate constants (<span class="html-italic">k</span>) of RhB; (<b>d</b>) comparison of the Co-C/SiO<sub>2</sub> catalyst with other Fenton-like catalysts. Conditions: [Catalyst] = 0.2 g/L, [PMS] = 0.2 g/L, [RhB] = 25 mg/L, pH = 6.6, T = 25 °C.</p>
Full article ">Figure 4
<p>(<b>a</b>) Universality of Co-C/SiO<sub>2</sub>; (<b>b</b>) the corresponding kinetics and (<b>c</b>) the corresponding degradation rate constants (<span class="html-italic">k</span>); (<b>d</b>) RhB removal with different scavengers; (<b>e</b>) the corresponding kinetics and (<b>f</b>) the corresponding degradation rate constants (<span class="html-italic">k</span>) of RhB. Conditions: [Catalyst] = 0.2 g/L, [PMS] = 0.2 g/L, [RhB] = 25 mg/L, pH = 6.6, T = 25 °C; (<b>g</b>) diagram reaction of DPA with <sup>1</sup>O<sub>2</sub>; (<b>h</b>) UV-vis absorption spectrum of DPA at different concentrations; (<b>i</b>) the fitting curve of concentration gradient of DPA at the absorption peak of 372 nm.</p>
Full article ">Figure 5
<p>(<b>a</b>) Catalytic activity of Co-C/SiO<sub>2</sub> for different cycle numbers. (<b>b</b>) XRD diffraction pattern of Co-C/SiO<sub>2</sub> before and after the reaction. (<b>c</b>) XPS of Co-C/SiO<sub>2</sub> before and after the reaction.</p>
Full article ">
24 pages, 5335 KiB  
Article
Biomass Waste Utilization as Nanocomposite Anodes through Conductive Polymers Strengthened SiO2/C from Streblus asper Leaves for Sustainable Energy Storages
by Thanapat Autthawong, Natthakan Ratsameetammajak, Kittiched Khunpakdee, Mitsutaka Haruta, Torranin Chairuangsri and Thapanee Sarakonsri
Polymers 2024, 16(10), 1414; https://doi.org/10.3390/polym16101414 - 16 May 2024
Viewed by 815
Abstract
Sustainable anode materials, including natural silica and biomass-derived carbon materials, are gaining increasing attention in emerging energy storage applications. In this research, we highlighted a silica/carbon (SiO2/C) derived from Streblus asper leaf wastes using a simple method. Dried Streblus asper leaves, [...] Read more.
Sustainable anode materials, including natural silica and biomass-derived carbon materials, are gaining increasing attention in emerging energy storage applications. In this research, we highlighted a silica/carbon (SiO2/C) derived from Streblus asper leaf wastes using a simple method. Dried Streblus asper leaves, which have plenty of biomass in Thailand, have a unique leaf texture due to their high SiO2 content. We can convert these worthless leaves into SiO2/C nanocomposites in one step, producing eco-materials with distinctive microstructures that influence electrochemical energy storage performance. Through nanostructured design, SiO2/C is thoroughly covered by a well-connected framework of conductive hybrid polymers based on the sodium alginate–polypyrrole (SA-PPy) network, exhibiting impressive morphology and performance. In addition, an excellent electrically conductive SA-PPy network binds to the SiO2/C particle surface through crosslinker bonding, creating a flexible porous space that effectively facilitates the SiO2 large volume expansion. At a current density of 0.3 C, this synthesized SA-PPy@Nano-SiO2/C anode provides a high specific capacity of 756 mAh g−1 over 350 cycles, accounting for 99.7% of the theoretical specific capacity. At the high current of 1 C (758 mA g−1), a superior sustained cycle life of over 500 cycles was evidenced, with over 93% capacity retention. The research also highlighted the potential for this approach to be scaled up for commercial production, which could have a significant impact on the sustainability of the lithium-ion battery industry. Overall, the development of green nanocomposites along with polymers having a distinctive structure is an exciting area of research that has the potential to address some of the key challenges associated with lithium-ion batteries, such as capacity degradation and safety concerns, while also promoting sustainability and reducing environmental impact. Full article
(This article belongs to the Special Issue Carbon/Polymer Composite Materials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD patterns and (<b>b</b>) FTIR spectra of synthesized products: Nano-SiO<sub>2</sub>, Nano-SiO<sub>2</sub>/C, PPy, and PPy@Nano-SiO<sub>2</sub>/C nanocomposites; (<b>c</b>) Raman spectra; and (<b>d</b>) thermogravimetric analysis curves under an air atmosphere for synthesized Nano-SiO<sub>2</sub>/C and PPy@Nano-SiO<sub>2</sub>/C nanocomposites.</p>
Full article ">Figure 2
<p>SEM images of the as-prepared products: (<b>a</b>) Nano-SiO<sub>2</sub>, (<b>b</b>) Nano-SiO<sub>2</sub>/C, and (<b>c</b>) PPy@Nano-SiO<sub>2</sub>/C nanocomposites; (<b>d</b>) the SEM-EDS mapping area corresponded to the elemental mapping of the PPy@Nano-SiO<sub>2</sub>/C nanocomposite: the corresponding elemental mapping of (<b>e</b>) C, (<b>f</b>) Si, (<b>g</b>) O, and (<b>h</b>) N, respectively; (<b>i</b>) schematic diagram of morphology in as-synthesized materials.</p>
Full article ">Figure 3
<p>CV curves of prepared electrodes at the first three cycles of the PPy@Nano-SiO<sub>2</sub>/C electrode between 0.01 and 3.0 V at a scan rate of 0.1 mV s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>TEM images corresponding SAED patterns (inset) of pre-cycled (<b>a</b>,<b>b</b>) and post-cycled (<b>d</b>,<b>e</b>) PPy@Nano-SiO<sub>2</sub>/C electrodes and particle size histograms of SiO<sub>2</sub> NPs (<b>c</b>) and Si QDs (<b>f</b>), HRTEM image of a single Si-QDs (<b>g</b>) showing the (111) lattice (inset), HAADF-STEM image (<b>h</b>), and EDS elemental mapping images of SA-PPy@Nano-SiO<sub>2</sub>/C: overlay of C K edge (<b>i</b>), O K edge (<b>j</b>), Si K edge (<b>k</b>), and N K edge (<b>l</b>).</p>
Full article ">Figure 5
<p>Battery performances of prepared electrodes: (<b>a</b>) rate cycle capability at different current densities in the range of 0.1C−1.0C, (<b>b</b>) comparative cycle performance and the corresponding Coulombic efficiency at a current density of 0.3C, (<b>c</b>) long-term cycle stability at a charging state of 1.0C for 500 cycles.</p>
Full article ">Figure 6
<p>(<b>a</b>) Nyquist plots with a fitted equivalent circuit inset of prepared electrodes, and (<b>b</b>) Warburg coefficient plots for the initial state.</p>
Full article ">Figure 7
<p>Dynamic analysis of PPy@Nano-SiO<sub>2</sub>/C electrode in a half-cell configuration: (<b>a</b>) CV curves recorded at different scan rates; (<b>b</b>) b-value of a relationship between the log (sweep rate, mV s<sup>−1</sup>) and log (peak current, mA) in the discharge and charge processes (marked as peak A and peak B in (<b>a</b>)), (<b>c</b>) CV curves at the scan rate of 2.0 mV s<sup>−1</sup> with capacitive-controlled (green region) contribution; and (<b>d</b>) variation of capacitive and diffusion contribution at different scan rates.</p>
Full article ">Figure 8
<p>(<b>a</b>) Visualization dissolution tests of PPy@Nano-SiO<sub>2</sub>/C electrode in 1 M LiPF<sub>6</sub> EC/DMC electrolyte, (<b>b</b>) Raman and (<b>c</b>) FTIR spectra of the collected electrolyte after being stored in the Ar-filled glovebox at 25 °C for 7 days.</p>
Full article ">Figure 9
<p>Schematic diagram of SA-PPy@Nano-SiO<sub>2</sub>/C electrode and adhesive polymer network of PPy-SA-PPy.</p>
Full article ">
11 pages, 2402 KiB  
Article
Influence of Silica Nanoparticles on the Physical Properties of Random Polypropylene
by Evangelia Delli, Dimitrios Gkiliopoulos, Evangelia Vouvoudi, Dimitrios N. Bikiaris, Thomas Kehagias and Konstantinos Chrissafis
J. Compos. Sci. 2024, 8(5), 186; https://doi.org/10.3390/jcs8050186 - 16 May 2024
Viewed by 724
Abstract
Random polypropylene is considered an alternative material to regular polypropylene for applications where improved impact and creep resistance, as well as stiffness, are required. Random polypropylene nanocomposites reinforced with dimethyldichlorosilane-treated silica particles were prepared using meltmixing. The effect of varying the nanoparticles’ content [...] Read more.
Random polypropylene is considered an alternative material to regular polypropylene for applications where improved impact and creep resistance, as well as stiffness, are required. Random polypropylene nanocomposites reinforced with dimethyldichlorosilane-treated silica particles were prepared using meltmixing. The effect of varying the nanoparticles’ content on the structural, mechanical, damping and thermal behavior of the nanocomposites was investigated. The results indicated the improved deformation potential, fracture toughness, and energy storage capacity of the matrix with increasing the filler content. It was observed that the use of high filler fractions limited the reinforcing efficiency of the SiO2 nanoparticles due to the formation of large agglomerates. The nanoparticles’ segregation was initially advised by modeling Young’s modulus but was also confirmed by electron imaging. Examination of the thermal properties of the nanocomposites indicated the limited effect of the nanoparticles on the melting behavior along with the thermal stability of the matrix. These results confirmed the usage of silica nanoparticles as a way of further improving the mechanical and thermomechanical properties of random polypropylene. Full article
(This article belongs to the Special Issue Feature Papers in Journal of Composites Science in 2024)
Show Figures

Figure 1

Figure 1
<p>Structural characterization: (<b>a</b>) FT-IR transmittance spectrum and (<b>b</b>) X-ray diffraction pattern of the SiO<sub>2</sub> NP composites, neat PP<sub>R</sub>, and SiO<sub>2</sub> nanoparticles.</p>
Full article ">Figure 2
<p>Thermal characterization: thermograms recorded during (<b>a</b>) DSC heating and (<b>b</b>) DSC cooling of PP<sub>R</sub> and its nanocomposites, while (<b>c</b>) TGA weight loss curves of all samples pyrolyzed and (<b>d</b>) the derived curves of the mass loss rate are also illustrated.</p>
Full article ">Figure 3
<p>Mechanical characterization: effect of SiO<sub>2</sub> NP content on (<b>a</b>) Young’s modulus of the nanocomposites. The data were fit using the Einstein model. (<b>b</b>) Tensile strength, (<b>c</b>) elongation at break results, and (<b>d</b>) impact toughness as a function of the nanoparticles content. The SiO<sub>2</sub> NP content values used are based on those obtained from the actual TGA analysis of each sample. The presented data correspond to the average values obtained out of five specimens tested for each composition, while error bars indicate the standard deviation.</p>
Full article ">Figure 4
<p>SEM imaging of the PP<sub>R</sub>/SiO<sub>2</sub>nanocomposites. Images of the fracture area of dumbbell-shaped specimens obtained for (<b>a</b>) neat PP<sub>R</sub> and its nanocomposites with (<b>b</b>) 1 wt.%, (<b>c</b>) 2.5 wt.%, (<b>d</b>) 5 wt.%, and (<b>e</b>) 10 wt.% SiO<sub>2</sub> NPs. The red arrows indicate the position of SiO<sub>2</sub> nanoparticle agglomerations.</p>
Full article ">Figure 5
<p>TEM imaging of the PP<sub>R</sub>/SiO<sub>2</sub> nanocomposites:bright field micrographs of (<b>a</b>) PP<sub>R</sub>/SiO<sub>2</sub> 1 wt.% and (<b>b</b>) PP<sub>R</sub>/SiO<sub>2</sub> 10 wt.% nanocomposites. The insetsshow magnified images of the nanoparticles’ aggregates in the polymeric matrix for the two cases, respectively.</p>
Full article ">Figure 6
<p>DMA characterization: (<b>a</b>) storage modulus <span class="html-italic">E</span>′ and (<b>b</b>) tan<span class="html-italic">δ</span> curves of neat PP<sub>R</sub> and the PP<sub>R</sub>/SiO<sub>2</sub> composites as a function of temperature.</p>
Full article ">
15 pages, 4317 KiB  
Article
Magnetic Aerogels for Room-Temperature Catalytic Production of Bis(indolyl)methane Derivatives
by Nicola Melis, Danilo Loche, Swapneel V. Thakkar, Maria Giorgia Cutrufello, Maria Franca Sini, Gianmarco Sedda, Luca Pilia, Angelo Frongia and Maria Francesca Casula
Molecules 2024, 29(10), 2223; https://doi.org/10.3390/molecules29102223 - 9 May 2024
Viewed by 662
Abstract
The potential of aerogels as catalysts for the synthesis of a relevant class of bis-heterocyclic compounds such as bis(indolyl)methanes was investigated. In particular, the studied catalyst was a nanocomposite aerogel based on nanocrystalline nickel ferrite (NiFe2O4) dispersed on amorphous [...] Read more.
The potential of aerogels as catalysts for the synthesis of a relevant class of bis-heterocyclic compounds such as bis(indolyl)methanes was investigated. In particular, the studied catalyst was a nanocomposite aerogel based on nanocrystalline nickel ferrite (NiFe2O4) dispersed on amorphous porous silica aerogel obtained by two-step sol–gel synthesis followed by gel drying under supercritical conditions and calcination treatments. It was found that the NiFe2O4/SiO2 aerogel is an active catalyst for the selected reaction, enabling high conversions at room temperature, and it proved to be active for three repeated runs. The catalytic activity can be ascribed to both the textural and acidic features of the silica matrix and of the nanocrystalline ferrite. In addition, ferrite nanocrystals provide functionality for magnetic recovery of the catalyst from the crude mixture, enabling time-effective separation from the reaction environment. Evidence of the retention of species involved in the reaction into the catalyst is also pointed out, likely due to the porosity of the aerogel together with the affinity of some species towards the silica matrix. Our work contributes to the study of aerogels as catalysts for organic reactions by demonstrating their potential as well as limitations for the room-temperature synthesis of bis(indolyl)methanes. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Mechanism for the formation of bis(indolyl)methane <b>3</b> from indole <b>1</b> and 4-nitrobenzaldehyde <b>2</b>. Red arrows suggest possible electron rearrangements.</p>
Full article ">Figure 2
<p>Schematic of the production procedure of the aerogel catalysts: (<b>a</b>) sol–gel synthesis of the multicomponent gel by co-hydrolysis and co-gelation of the metal and silica precursors; (<b>b</b>) aerogel production by high-temperature supercritical drying of the multicomponent gel; (<b>c</b>) thermal treatments to promote the formation of magnetic NiFe<sub>2</sub>O<sub>4</sub>/SiO<sub>2</sub> nanocomposite aerogel catalysts.</p>
Full article ">Figure 3
<p>Representative images of the aerogel catalyst at different stages of nanocomposite preparation: (<b>a</b>) optical image of a highly porous nickel-containing composite aerogel as obtained after supercritical drying; (<b>b</b>) corresponding SEM image and (<b>c</b>–<b>f</b>) energy-filtered images showing oxygen distribution (<b>c</b>), silicon distribution (<b>d</b>), Fe distribution (<b>e</b>), and nickel distribution (<b>f</b>). TEM images (scale bar is 100 nm) of the NiFe<sub>2</sub>O<sub>4</sub>/SiO<sub>2</sub> aerogel catalyst as obtained by calcination at 900 °C (<b>g</b>,<b>h</b>).</p>
Full article ">Figure 4
<p><sup>1</sup>H NMR spectra of the reaction mixture as obtained without catalyst (red bottom curve); with the use of plain SiO<sub>2</sub> aerogel catalyst (green intermediate curve); and with the use of NiFeO<sub>2</sub>-SiO<sub>2</sub> aerogel catalyst (top blue curve). Significant spectral ranges with corresponding attribution are included as a guide (catalyst amount: 5 mol %; run time: 1 week; solvent: CH<sub>2</sub>Cl<sub>2</sub>; reaction temperature: RT).</p>
Full article ">Figure 5
<p>BIMs synthesis catalyzed by Ni-CAT aerogel catalysts (catalyst amount: 5 mol %; run time: 1 week; solvent: CH<sub>2</sub>Cl<sub>2</sub>; reaction temperature: RT): the composition of the resulting reaction mixture is represented as relative amounts of reactant <b>2</b> (grey bars) and product <b>3</b> (brown bars).</p>
Full article ">Figure 6
<p><sup>1</sup>H-NMR spectra of pure BIM (green line), and the extracts after 1 (blue line) and 3 (red line) catalytic runs.</p>
Full article ">
14 pages, 5350 KiB  
Article
The Optimization of the One-Pot Synthesis of Au@SiO2 Core–Shell Nanostructures: Modification with Dansyl Group and Their Fluorescent Properties
by Agata Kowalska, Elżbieta Adamska, Anna Synak and Beata Grobelna
Materials 2024, 17(10), 2213; https://doi.org/10.3390/ma17102213 - 8 May 2024
Viewed by 1291
Abstract
This work describes the optimization of the one-pot synthesis of fine core–shell nanostructures based on nanogold (Au NPs) and silica (SiO2). The obtained core–shell nanomaterials were characterized by Transmission Electron Microscopy (TEM and by the method of spectroscopes such as UV–Vis [...] Read more.
This work describes the optimization of the one-pot synthesis of fine core–shell nanostructures based on nanogold (Au NPs) and silica (SiO2). The obtained core–shell nanomaterials were characterized by Transmission Electron Microscopy (TEM and by the method of spectroscopes such as UV–Vis Spectroscopy and Fourier Transform Infrared Spectroscopy (FT-IR). In addition, the measurement of the zeta potential and size of the obtained particles helped present a full characterization of Au@SiO2 nanostructures. The results show that the influence of reagents acting as reducers, stabilizers, or precursors of the silica shell affects the morphology of the obtained material. By controlling the effect of the added silica precursor, the thickness of the shell can be manipulated, the reducer has an effect on the shape and variety, and then the stabilizer affects their agglomeration. This work provides also a new approach for Au@SiO2 core–shell nanostructure preparation by further modification with dansyl chloride (DNS–Cl). The results show that, by tuning the silica shell thickness, the intensity of the fluorescence spectrum of Au@SiO2–(CH2)3–NH–DNS nanocomposite is about 12 times higher than that of DNS–Cl. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The synthesis route of the presented Au@SiO<sub>2</sub> nanostructures.</p>
Full article ">Figure 2
<p>Schematic route to modification and functionalization of Au@SiO<sub>2</sub> nanostructure.</p>
Full article ">Figure 3
<p>TEM images for (<b>a</b>) Au NPs and (<b>b</b>) SiO<sub>2</sub> NPs.</p>
Full article ">Figure 4
<p>Absorption spectra for (<b>a</b>) Au NPs and SiO<sub>2</sub> NPs and (<b>b</b>) all samples of Au@SiO<sub>2</sub> nanocomposites.</p>
Full article ">Figure 5
<p>FT-IR spectra for (<b>a</b>) SiO<sub>2</sub> NPs and (<b>b</b>) all samples of Au@SiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>Graph for zeta potential distribution of the synthesized samples of Au@SiO<sub>2</sub>.</p>
Full article ">Figure 7
<p>Graph for particle diameter of the synthesized samples of Au@SiO<sub>2</sub>.</p>
Full article ">Figure 8
<p>TEM images of syntheses to compare the effect of TEOS concentration: (<b>a</b>) 4.38 × 10<sup>−3</sup> M (1st synthesis), (<b>b</b>) 2.8 × 10<sup>−3</sup> M (2nd synthesis) and (<b>c</b>) 2.33 × 10<sup>−3</sup> M (6th synthesis).</p>
Full article ">Figure 9
<p>TEM images of syntheses to compare the influence of CTAB concentration: (<b>a</b>) 7.47 × 10<sup>−4</sup> M (Synthesis 1), (<b>b</b>) 1.07 × 10<sup>−3</sup> M (Synthesis 3), and (<b>c</b>) 1.6 × 10<sup>−3</sup> M (Synthesis 5).</p>
Full article ">Figure 10
<p>TEM images of syntheses to compare the influence of hydrazine concentration: (<b>a</b>) 4.24 × 10<sup>−1</sup> M (Synthesis 6), (<b>b</b>) 3.03 × 10<sup>−1</sup> M (Synthesis 1), (<b>c</b>) 2.43 × 10<sup>−1</sup> M (Synthesis 3), and (<b>d</b>) 1.82 × 10<sup>−1</sup> M (Synthesis 4).</p>
Full article ">Figure 11
<p>Absorption spectra of Au@SiO<sub>2</sub>–(CH<sub>2</sub>)<sub>3</sub>–NH–Gly–OH.</p>
Full article ">Figure 12
<p>(<b>a</b>) Normalized absorption (blue line) and emission spectra (green line) of Au@SiO<sub>2</sub>–(CH<sub>2</sub>)<sub>3</sub>–NH–DNS in methanol solution. (<b>b</b>) Fluorescence spectra of Au@SiO<sub>2</sub>–(CH<sub>2</sub>)<sub>3</sub>–NH–DNS and DNS–Cl in methanol.</p>
Full article ">
Back to TopTop