[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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (9,867)

Search Parameters:
Keywords = graphene

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 8704 KiB  
Article
Immunomodulatory R848-Loaded Anti-PD-L1-Conjugated Reduced Graphene Oxide Quantum Dots for Photothermal Immunotherapy of Glioblastoma
by Yu-Jen Lu, Reesha Kakkadavath Vayalakkara, Banendu Sunder Dash, Shang-Hsiu Hu, Thejas Pandaraparambil Premji, Chun-Yuan Wu, Yang-Jin Shen and Jyh-Ping Chen
Pharmaceutics 2024, 16(8), 1064; https://doi.org/10.3390/pharmaceutics16081064 (registering DOI) - 13 Aug 2024
Abstract
Glioblastoma multiforme (GBM) is the most severe form of brain cancer and presents unique challenges to developing novel treatments due to its immunosuppressive milieu where receptors like programmed death ligand 1 (PD-L1) are frequently elevated to prevent an effective anti-tumor immune response. To [...] Read more.
Glioblastoma multiforme (GBM) is the most severe form of brain cancer and presents unique challenges to developing novel treatments due to its immunosuppressive milieu where receptors like programmed death ligand 1 (PD-L1) are frequently elevated to prevent an effective anti-tumor immune response. To potentially shift the GBM environment from being immunosuppressive to immune-enhancing, we engineered a novel nanovehicle from reduced graphene oxide quantum dot (rGOQD), which are loaded with the immunomodulatory drug resiquimod (R848) and conjugated with an anti-PD-L1 antibody (aPD-L1). The immunomodulatory rGOQD/R8/aPDL1 nanoparticles can actively target the PD-L1 on the surface of ALTS1C1 murine glioblastoma cells and release R848 to enhance the T-cell-driven anti-tumor response. From in vitro experiments, the PD-L1-mediated intracellular uptake and the rGOQD-induced photothermal response after irradiation with near-infrared laser light led to the death of cancer cells and the release of damage-associated molecular patterns (DAMPs). The combinational effect of R848 and released DAMPs synergistically produces antigens to activate dendritic cells, which can prime T lymphocytes to infiltrate the tumor in vivo. As a result, T cells effectively target and attack the PD-L1-suppressed glioma cells and foster a robust photothermal therapy elicited anti-tumor immune response from a syngeneic mouse model of GBM with subcutaneously implanted ALTS1C1 cells. Full article
(This article belongs to the Special Issue Metal and Carbon Nanomaterials for Pharmaceutical Applications)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic illustration of the preparation of rGOQD/R8/aPDL1, including reduction of GOQD to rGOQD using polyethylene (PEI), immunoadjuvant drug R848 loading on rGOQD by π–π stacking (rGOQD/R8), and aPD-L1 conjugation on rGO/R8 using amine groups in rGOQD and aldehyde group in activated aPD-L1. (<b>b</b>) The photo-immunotherapy using rGOQD/R8/aPDL1 involves photothermal therapy and immune cell activation to exert an anti-tumor effect by (1) binding to the overexpressed PD-L1 receptors on tumor cell surface; (2) R484 release for activation of adaptive immune response; (3) photothermal-effect-induced cell death; (4) antigen release and antigen-presenting cells (APCs) activation; (5) dendritic cells (DCs) activation; (6) T cells recruitment.</p>
Full article ">Figure 2
<p>(<b>a</b>) The TEM image of rGOQD/R8/aPDL1 (scale bar = 500 nm). (<b>b</b>) The size distribution from dynamic light scattering analysis of GOQD, rGOQD, rGOQD/R8, and rGOQD/R8/aPDL1. (<b>c</b>) The zeta potential values of different nanoparticles (mean ± SD, <span class="html-italic">n</span> = 3). (<b>d</b>) The FTIR spectra of GOQD, rGOQD, and rGOQD/R8. (<b>e</b>) The UV-Vis spectroscopy analysis of GOQD, rGOQD, rGOQD/R8, and rGOQD/R8/aPDL1.</p>
Full article ">Figure 3
<p>The photothermal images (<b>a</b>), and the corresponding temperature profiles (<b>b</b>) by irradiating GOQD or rGOQD/R8 (100 μg/mL) with 808 nm laser (1.5 W/cm<sup>2</sup>) for 5 min. The control is deionized water (DIW). The thermal images (<b>c</b>), and the corresponding temperature profiles (<b>d</b>) by irradiating 25–100 μg/mL rGOQD/R8/aPDL1 with 808 nm laser (1.5 W/cm<sup>2</sup>) for 5 min. (<b>e</b>) The in vitro release of R848 from rGOQD/R8/aPDL1 at pH 5 and 7.4. (<b>f</b>) The stability of rGOQD/R8/aPDL1 in PBS and DMEM cell culture medium by measuring the particle size from DLS. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>The biocompatibility of rGOQD/R8/aPDL1 was tested with 3T3 mouse embryonic fibroblast cells (<b>a</b>) and ALTS1C1 mouse glioma cells (<b>b</b>) with MTS assays at 24, 48, and 72 h. (<b>c</b>) The flow cytometry analysis of intracellular uptake of Cy5.5-tagged rGOQD/R8 and rGOQD/R8/aPDL1 after incubation with ALTS1C1 cells for 4 and 24 h. (<b>d</b>) The corresponding quantified fluorescence intensity from flow cytometry analysis of intracellular uptake of Cy5.5-tagged nanoparticles. * <span class="html-italic">p</span> &lt; 0.05 compared with rGOQD/R8. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 5
<p>(<b>a</b>) The cell viability of ALTS1C1 cells after incubating cells with rGOQD/R8 (rGOQD/R8+L) or rGOQD/R8/aPDL1 (rGOQD/R8/aPDL1+L) at varying concentrations and irradiated with 808 nm laser at 1.5 W/cm<sup>2</sup> for 5 min. (<b>b</b>) The fluorescence microscopy images of Calcein-AM and PI co-stained cells after incubation with different nanoparticles and 808 nm NIR treatment at 1.5 W/cm<sup>2</sup> for 5 min. Cells treated with PBS and rGOQD/R8 without laser irradiation were used as controls. (<b>c</b>) The CLSM images of ALTS1C1 cells by staining with fluorescein-tagged anti-calreticulin antibody after different treatments. The rGOQD/R8+L and rGOQD/R8/aPDL1+L groups are irradiated with 808 nm NIR laser at 1.5 W/cm<sup>2</sup> for 5 min. (<b>d</b>) The corresponding quantification of calreticulin fluorescence intensity by using the PAX-it software. All data are represented as mean ± SD (<span class="html-italic">n</span> = 3). <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">Figure 6
<p>The ex vivo fluorescence images of major organs (<b>a</b>) and the corresponding quantification of nanoparticle distribution in each organ from fluorescence intensity (<b>b</b>) with an in vivo imaging system (IVIS) 4 h after administration of Cy5.5-labelled rGOQD/R8/aPDL1 or rGOQD/R8 to ALTS1C1 tumor-bearing mice through the tail vein. The ex vivo fluorescence images (<b>c</b>) and the corresponding fluorescence intensity (<b>d</b>) of tumors with an IVIS 4 h after administration of Cy5.5-labelled rGOQD/R8/aPDL1 or rGOQD/R8 to ALTS1C1 tumor-bearing mice through the tail vein.</p>
Full article ">Figure 7
<p>The thermal images (<b>a</b>) and the corresponding in vivo peak temperature profiles (<b>b</b>) of ALTS1C1 tumor-bearing mice after intravenous injection of rGOQD/R8/aPDL1 or rGOQD/R8 followed by 808 nm NIR laser irradiation 24 h post injection (mean ± SD, <span class="html-italic">n</span> = 3). The activation of dendritic cells by rGOQD/R8/aPDL1 or rGOQD/R8 was compared with confocal images of lymph nodes 19 days after treatment by immunofluorescence staining of CD11C (antigen-presenting cells) and CD86 (dendritic cells) in red and green, respectively, and counterstaining the nucleus with DAPI in blue (scale bar = 50 μm) (<b>c</b>).</p>
Full article ">Figure 8
<p>The in vivo therapeutic evaluation was studied using a syngeneic mouse model of GBM with subcutaneously implanted ALTSC1 cells. The mice were divided into four groups and the treatment was initiated by injection of samples on day 10, followed by intravenous injection on days 13, 17, and 20. The control group is PBS and the rGOQD/R8+L and rGOQD/R8/PDL1 groups are with 808 nm NIR laser irradiation at 1.5 W/cm<sup>2</sup> for 5 min. The tumor volume change (<b>a</b>), the scattered plot of tumor volume on day 21 (<b>b</b>), and the survival curve of animals (<b>c</b>) of ALTSC1 tumor-bearing mice after different treatments (mean ± SD, <span class="html-italic">n</span> = 3). The sacrificing criteria were when the tumor volume exceeded 1000 mm<sup>3</sup>. <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">Figure 9
<p>(<b>a</b>) In vivo immune response after treatment with rGOQD/R8, rGOQD/R8+L, and rGOQD/R8/aPDL1+L. The infiltration of T cells into the tumor after various treatments was measured by immunofluorescence staining with anti-CD4 and anti-CD8 antibodies. The confocal immunofluorescence images and the corresponding quantification of fluorescence intensity of CD4 and CD8 in the tumor tissues 19 days after treatments (scale bar = 50 μm). (<b>b</b>) The immunohistochemistry (IHC) of PD-L1, tumor necrosis factor (TNF-α), and Ki-67, and H&amp;E staining of tumor tissues 19 days after treatments. <sup>α</sup> <span class="html-italic">p</span> &lt; 0.05 compared to PBS, <sup>β</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8, <sup>γ</sup> <span class="html-italic">p</span> &lt; 0.05 compared to rGOQD/R8+L.</p>
Full article ">
8 pages, 1729 KiB  
Communication
H2 Adsorption on Small Pd-Ni Clusters Deposited on N-Doped Graphene: A Theoretical Study
by Brenda García-Hilerio, Lidia Santiago-Silva, Adriana Vásquez-García, Alejandro Gomez-Sanchez, Víctor A. Franco-Luján and Heriberto Cruz-Martínez
C 2024, 10(3), 73; https://doi.org/10.3390/c10030073 (registering DOI) - 13 Aug 2024
Abstract
The study of novel materials for H2 storage is essential to consolidate the hydrogen as a clean energy source. In this sense, the H2 adsorption on Pd4-nNin (n = 0–3) clusters embedded on pyridinic-type N-doped graphene (PNG) was [...] Read more.
The study of novel materials for H2 storage is essential to consolidate the hydrogen as a clean energy source. In this sense, the H2 adsorption on Pd4-nNin (n = 0–3) clusters embedded on pyridinic-type N-doped graphene (PNG) was investigated using density functional theory calculations. First, the properties of Pd4-nNin (n = 0–3) clusters embedded on PNG were analyzed in detail. Then, the H2 adsorption on these composites was computed. The Eint between the Pd4-nNin (n = 0–3) clusters and the PNG was greater than that computed in the literature for Pd-based systems embedded on pristine graphene. Consequently, it was deduced that PNG can more significantly stabilize the Pd4-nNin (n = 0–3) clusters. The analyzed composites exhibited a HOMO–LUMO gap less than 1 eV, indicating good reactivity. Based on the Eads of H2 on Pd4-nNin (n = 0–3) clusters embedded on PNG, it was observed that the analyzed systems meet the standards set by the DOE. Therefore, these composites can be viable alternatives for hydrogen storage. Full article
(This article belongs to the Special Issue Adsorption on Carbon-Based Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The PNG structure. Blue, yellow, and white spheres represent N, C, and H atoms, respectively.</p>
Full article ">Figure 2
<p>The most stable interactions between the Pd<sub>4-n</sub>Ni<sub>n</sub> (n = 0–3) clusters and the PNG. (<b>a</b>) Pd<sub>4</sub> cluster embedded on PNG, (<b>b</b>) Pd<sub>3</sub>Ni<sub>1</sub> cluster embedded on PNG, (<b>c</b>) Pd<sub>2</sub>Ni<sub>2</sub> cluster embedded on PNG, and (<b>d</b>) Pd<sub>1</sub>Ni<sub>3</sub> cluster embedded on PNG. Blue, yellow, white, green, and black spheres represent N, C, H, Ni, and Pd atoms, respectively.</p>
Full article ">Figure 3
<p>Spin density (red) plots of the Pd<sub>4-n</sub>Ni<sub>n</sub> (n = 0–3) clusters embedded on PNG. (<b>a</b>) Pd<sub>4</sub> cluster embedded on PNG, (<b>b</b>) Pd<sub>3</sub>Ni<sub>1</sub> cluster embedded on PNG, (<b>c</b>) Pd<sub>2</sub>Ni<sub>2</sub> cluster embedded on PNG, and (<b>d</b>) Pd<sub>1</sub>Ni<sub>3</sub> cluster embedded on PNG. Blue, yellow, white, green, and black spheres represent N, C, H, Ni, and Pd atoms, respectively.</p>
Full article ">Figure 4
<p>The BCPs (orange spheres) and bond paths between the Pd<sub>4-n</sub>Ni<sub>n</sub> (n = 0–3) clusters and the PNG. (<b>a</b>) Pd<sub>4</sub> cluster embedded on PNG, (<b>b</b>) Pd<sub>3</sub>Ni<sub>1</sub> cluster embedded on PNG, (<b>c</b>) Pd<sub>2</sub>Ni<sub>2</sub> cluster embedded on PNG, and (<b>d</b>) Pd<sub>1</sub>Ni<sub>3</sub> cluster embedded on PNG. Blue, yellow, white, green, and black spheres represent N, C, H, Ni, and Pd atoms, respectively.</p>
Full article ">Figure 5
<p>The most stable H<sub>2</sub> adsorptions on Pd<sub>4-n</sub>Ni<sub>n</sub> (n = 0–3) clusters embedded on PNG. (<b>a</b>) H<sub>2</sub> adsorption on Pd<sub>4</sub> cluster embedded on PNG, (<b>b</b>) H<sub>2</sub> adsorption on Pd<sub>3</sub>Ni<sub>1</sub> cluster embedded on PNG, (<b>c</b>) H<sub>2</sub> adsorption on Pd<sub>2</sub>Ni<sub>2</sub> cluster embedded on PNG, and (<b>d</b>) H<sub>2</sub> adsorption on Pd<sub>1</sub>Ni<sub>3</sub> cluster embedded on PNG. Blue, yellow, white, green, and black spheres represent N, C, H, Ni, and Pd atoms, respectively.</p>
Full article ">
17 pages, 10250 KiB  
Article
Planar Micro-Supercapacitors with High Power Density Screen-Printed by Aqueous Graphene Conductive Ink
by Youchang Wang, Xiaojing Zhang, Yuwei Zhu, Xiaolu Li and Zhigang Shen
Materials 2024, 17(16), 4021; https://doi.org/10.3390/ma17164021 (registering DOI) - 13 Aug 2024
Abstract
Simple and scalable production of micro-supercapacitors (MSCs) is crucial to address the energy requirements of miniature electronics. Although significant advancements have been achieved in fabricating MSCs through solution-based printing techniques, the realization of high-performance MSCs remains a challenge. In this paper, graphene-based MSCs [...] Read more.
Simple and scalable production of micro-supercapacitors (MSCs) is crucial to address the energy requirements of miniature electronics. Although significant advancements have been achieved in fabricating MSCs through solution-based printing techniques, the realization of high-performance MSCs remains a challenge. In this paper, graphene-based MSCs with a high power density were prepared through screen printing of aqueous conductive inks with appropriate rheological properties. High electrical conductivity (2.04 × 104 S∙m−1) and low equivalent series resistance (46.7 Ω) benefiting from the dense conductive network consisting of the mesoporous structure formed by graphene with carbon black dispersed as linkers, as well as the narrow finger width and interspace (200 µm) originating from the excellent printability, prompted the fully printed MSCs to deliver high capacitance (9.15 mF∙cm−2), energy density (1.30 µWh∙cm−2) and ultrahigh power density (89.9 mW∙cm−2). Notably, the resulting MSCs can effectively operate at scan rates up to 200 V∙s−1, which surpasses conventional supercapacitors by two orders of magnitude. In addition, the MSCs demonstrate excellent cycling stability (91.6% capacity retention and ~100% Coulombic efficiency after 10,000 cycles) and extraordinary mechanical properties (92.2% capacity retention after 5000 bending cycles), indicating their broad application prospects in flexible wearable/portable electronic systems. Full article
(This article belongs to the Section Carbon Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic synthesis of aqueous graphene conductive ink and screen-printed preparation of MSCs.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM image of pristine graphite before exfoliation. (<b>b</b>) TEM image of several stacked graphene nanosheets. (<b>c</b>) A typical AFM image of graphene. (<b>d</b>) SEM image of graphene after exfoliation. (<b>e</b>) HRTEM image of the rectangle region in (<b>b</b>). Histograms of the (<b>f</b>) thickness distribution and (<b>g</b>) area distribution of graphene.</p>
Full article ">Figure 3
<p>(<b>a</b>) Raman spectra and (<b>b</b>) C1s XPS of pristine graphite and graphene.</p>
Full article ">Figure 4
<p>(<b>a</b>) Viscosity of ink as a function of shear rate; (<b>b</b>) thixotropy of ink; (<b>c</b>) percentage recovery of ink viscosity; (<b>d</b>) variation of storage modulus (G′), and loss modulus (G″) with shear stress, where the solid and hollow symbols stand for G′ and G″, respectively; (<b>e</b>) loss coefficient tan δ as a function of shear stress; (<b>f</b>) optical microscope images of thin lines of ink printed on PET substrate by passing through the opening of a 100 μm optical microscope images of fine lines printed by ink through screen openings on PET substrates, from top to bottom corresponding to Ink–0.25, Ink–0.67, Ink–1.5, and Ink–4, respectively; (<b>g</b>) electrical conductivity of the printed patterns before and after hot pressing; (<b>h</b>) change in the relative electrical resistance of the printed patterns during 1000 repetitive bending cycles at a bending angle of 120°; (<b>i</b>) top-view SEM image of the fine line printed with Ink–1.5.</p>
Full article ">Figure 5
<p>(<b>a</b>) A representative cross-sectional SEM image of MSC. (<b>b</b>) A digital photograph of MSC<sub>1000</sub>, MSC<sub>500</sub>, and MSC<sub>200</sub>.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>f</b>) CV curves of MSC<sub>200</sub> at different scan rates of 0.01, 0.1, 1, 10, 100, and 200 V∙s<sup>−1</sup>. (<b>g</b>) The relationship of discharge current density varies with the scan rate for MSC<sub>200</sub>. (<b>h</b>) Nyquist plot of MSC<sub>200</sub>, MSC<sub>500</sub>, and MSC<sub>1000</sub>. (<b>i</b>) The relationship of impedance phase angle varies with frequency for the three MSCs.</p>
Full article ">Figure 7
<p>GCD curves at varying current densities of (<b>a</b>) 0.05–0.4 mA∙cm<sup>−2</sup> and (<b>b</b>) 0.5–1.0 mA∙cm<sup>−2</sup>. (<b>c</b>) Area capacitance of the three MSCs varying current density of 0.05–1.0 mA∙cm<sup>−2</sup>. (<b>d</b>) Comparison of area-specific capacitance of this study and other reported printed MSCs. (<b>e</b>) Cycling performance of MSC<sub>200</sub> at the current density of 0.5 mA∙cm<sup>−2</sup>. (<b>f</b>) Ragone plot on energy density and power density. (<b>g</b>) Photographs of MSC<sub>200</sub> at various bending angles. (<b>h</b>) Capacitance retention of MSC<sub>200</sub> during 5000 bending cycles. Inset: CV curves obtained at various bending angles using a scan rate of 200 mV∙s<sup>−1</sup> [<a href="#B15-materials-17-04021" class="html-bibr">15</a>,<a href="#B17-materials-17-04021" class="html-bibr">17</a>,<a href="#B18-materials-17-04021" class="html-bibr">18</a>,<a href="#B20-materials-17-04021" class="html-bibr">20</a>,<a href="#B22-materials-17-04021" class="html-bibr">22</a>,<a href="#B52-materials-17-04021" class="html-bibr">52</a>,<a href="#B56-materials-17-04021" class="html-bibr">56</a>,<a href="#B58-materials-17-04021" class="html-bibr">58</a>,<a href="#B59-materials-17-04021" class="html-bibr">59</a>,<a href="#B61-materials-17-04021" class="html-bibr">61</a>,<a href="#B62-materials-17-04021" class="html-bibr">62</a>].</p>
Full article ">
12 pages, 4616 KiB  
Article
Hot Deformation Behavior and Microstructure Evolution of a Graphene/Copper Composite
by Tiejun Li, Ruiyu Lu, Yuankui Cao, Bicheng Liu, Ao Fu and Bin Liu
Materials 2024, 17(16), 4010; https://doi.org/10.3390/ma17164010 - 12 Aug 2024
Viewed by 268
Abstract
Graphene/copper composites are promising in electronic and energy fields due to their superior conductivity, but microstructure control during thermal mechanical processing (TMP) remains a crucial issue for the manufacturing of high-performance graphene/copper composites. In this study, the hot deformation behavior of graphene/copper composites [...] Read more.
Graphene/copper composites are promising in electronic and energy fields due to their superior conductivity, but microstructure control during thermal mechanical processing (TMP) remains a crucial issue for the manufacturing of high-performance graphene/copper composites. In this study, the hot deformation behavior of graphene/copper composites was investigated by isothermal compression tests at deformation temperatures of 700~850 °C and strain rates of 0.01~10 s−1, and a constitutive equation based on the Arrhenius model and hot processing map was established. Results demonstrate that the deformation mechanism of the graphene/copper composites mainly involves dynamic recrystallization (DRX), and such DRX-mediated deformation behavior can be accurately described by the established Arrhenius model. In addition, it was found that the strain rate has a stronger impact on the DRX grain size than the deformation temperature. The optimum deformation temperature and strain rate were determined to be 800 °C and 1 s−1, respectively, with which a uniform microstructure with fine grains can be obtained. Full article
Show Figures

Figure 1

Figure 1
<p>Graphene/copper powders: (<b>a</b>) powder characteristics; (<b>b</b>) Raman spectroscopy.</p>
Full article ">Figure 2
<p>EBSD structure of graphene/copper composites sintered at 900 °C: (<b>a</b>) IQ map; (<b>b</b>) IPF map; (<b>c</b>) KAM map; (<b>d</b>) recrystallization map.</p>
Full article ">Figure 3
<p>Distribution of graphene in as-sintered composites: (<b>a</b>) low magnification; (<b>b</b>) high magnification. The inset shows the Raman spectrum of graphene.</p>
Full article ">Figure 4
<p>Macrophotographs of graphene/copper composites after hot deformation (black parts on the surface of the sample are oxidized skins that have not been completely cleaned off after hot deformation).</p>
Full article ">Figure 5
<p>True stress–strain curves of graphene/copper composites in different deformation conditions: (<b>a</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 10 s<sup>−1</sup>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 1 s<sup>−1</sup>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 0.1 s<sup>−1</sup>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo>˙</mo> </mover> </mrow> </semantics></math> = 0.01 s<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Fitting curve between TMP parameters and stress: (<b>a</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </mrow> <mo>−</mo> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mi>σ</mi> </mrow> </mrow> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </mrow> </mrow> </semantics></math>−<math display="inline"><semantics> <mrow> <mi>σ</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </mrow> </mrow> </semantics></math>−<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mrow> <mi>sin</mi> <mi mathvariant="normal">h</mi> </mrow> <mfenced separators="|"> <mrow> <mi mathvariant="sans-serif">α</mi> <mi>σ</mi> </mrow> </mfenced> </mrow> </mfenced> </mrow> </mrow> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mrow> <mi>sin</mi> <mi mathvariant="normal">h</mi> </mrow> <mfenced separators="|"> <mrow> <mi mathvariant="sans-serif">α</mi> <mi>σ</mi> </mrow> </mfenced> </mrow> </mfenced> </mrow> </mrow> </mrow> </semantics></math>−<math display="inline"><semantics> <mrow> <mfenced separators="|"> <mrow> <mrow> <mn>1</mn> <mo>/</mo> </mrow> <mi mathvariant="normal">T</mi> </mrow> </mfenced> </mrow> </semantics></math>; (<b>e</b>) <math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mi mathvariant="normal">Z</mi> </mrow> </mrow> </mrow> </semantics></math>−<math display="inline"><semantics> <mrow> <mrow> <mrow> <mi>ln</mi> </mrow> <mo>⁡</mo> <mrow> <mfenced open="[" close="]" separators="|"> <mrow> <mrow> <mi>sin</mi> <mi mathvariant="normal">h</mi> </mrow> <mfenced separators="|"> <mrow> <mi mathvariant="sans-serif">α</mi> <mi>σ</mi> </mrow> </mfenced> </mrow> </mfenced> </mrow> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 7
<p>Hot processing map for graphene/copper composites.</p>
Full article ">Figure 8
<p>IPF maps of graphene/copper composites at different deformation temperatures: (<b>a</b>) 750 °C; (<b>b</b>) 800 °C; (<b>c</b>) 850 °C.</p>
Full article ">Figure 9
<p>IPF diagrams of graphene/copper composites at different strain rates: (<b>a</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </semantics></math> = 0.1 s<sup>−1</sup>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </semantics></math> = 1 s<sup>−1</sup>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo mathvariant="normal">˙</mo> </mover> </mrow> </semantics></math> = 10 s<sup>−1</sup>.</p>
Full article ">
26 pages, 9190 KiB  
Review
The Recent Advancement of Graphene-Based Cathode Material for Rechargeable Zinc–Air Batteries
by Abrham Sendek Belete, Ababay Ketema Worku, Delele Worku Ayele, Addisu Alemayehu Assegie and Minbale Admas Teshager
Processes 2024, 12(8), 1684; https://doi.org/10.3390/pr12081684 - 12 Aug 2024
Viewed by 257
Abstract
Graphene-based materials (GBMs) are a prospective material of choice for rechargeable battery electrodes because of their unique set of qualities, which include tunable interlayer channels, high specific surface area, and strong electrical conductivity characteristics. The market for commercial rechargeable batteries is now dominated [...] Read more.
Graphene-based materials (GBMs) are a prospective material of choice for rechargeable battery electrodes because of their unique set of qualities, which include tunable interlayer channels, high specific surface area, and strong electrical conductivity characteristics. The market for commercial rechargeable batteries is now dominated by lithium-ion batteries (LIBs). One of the primary factors impeding the development of new energy vehicles and large-scale energy storage applications is the safety of LIBs. Zinc-based rechargeable batteries have emerged as a viable substitute for rechargeable batteries due to their affordability, safety, and improved performance. This review article explores recent developments in the synthesis and advancement of GBMs for rechargeable zinc–air batteries (ZABs) and common graphene-based electrocatalyst types. An outlook on the difficulties and probable future paths of this extremely promising field of study is provided at the end. Full article
(This article belongs to the Section Energy Systems)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Nominal cell voltages, volumetric energy densities, theoretical specific energies, and properties for various metal anodes; (<b>b</b>) schematic diagram of a ZAB; and (<b>c</b>) a comparison of the theoretical specific energies reversibility, stability, safety in aqueous media, and affordability of metal–air batteries. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B17-processes-12-01684" class="html-bibr">17</a>]. Copyright (2023), Springer, Berlin/Heidelberg, Germany.</p>
Full article ">Figure 2
<p>Schematic representation of prismatic zinc–air battery configuration. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B21-processes-12-01684" class="html-bibr">21</a>]. Copyright (2023), MDPI.</p>
Full article ">Figure 3
<p>Recent advances of graphene and its applications. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B54-processes-12-01684" class="html-bibr">54</a>]. Copyright (2022), MDPI.</p>
Full article ">Figure 4
<p>Diagram showing how graphene oxide nanoparticles are synthesized. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B60-processes-12-01684" class="html-bibr">60</a>]. Copyright (2022), Frontiers.</p>
Full article ">Figure 5
<p>Synthesis of graphene oxide using Hummer’s method. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B66-processes-12-01684" class="html-bibr">66</a>]. Copyright (2021), Nature.</p>
Full article ">Figure 6
<p>Schematic representation of the rapid synthesis of graphene oxide from graphene sheets. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B67-processes-12-01684" class="html-bibr">67</a>]. Copyright (2021), MDPI.</p>
Full article ">Figure 7
<p>Schematic representation of reduction of GO to RGO using chemical methods. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B75-processes-12-01684" class="html-bibr">75</a>]. Copyright (2017), RSC.</p>
Full article ">Figure 8
<p>Schematic illustration of the reduction of graphene oxide (GO) to reduced graphene oxide (RGO). Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B78-processes-12-01684" class="html-bibr">78</a>]. Copyright (2018), IOP.</p>
Full article ">Figure 9
<p>Schematic representation of the most frequently used synthesis techniques for graphene.</p>
Full article ">Figure 10
<p>Graphene and its derivatives. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B102-processes-12-01684" class="html-bibr">102</a>]. Copyright (2023), MDPI.</p>
Full article ">Figure 11
<p>STEM images of boron-doped graphene quantum structures (BGQS) (<b>a</b>) and TEM images of (<b>b</b>) as-prepared MoS<sub>2</sub>, (<b>c</b>) BGQS-10/MoS<sub>2</sub>, (<b>d</b>) BGQS-30/MoS<sub>2</sub>, (<b>e</b>) BGQS-50/MoS<sub>2</sub>, and (<b>f</b>) BGQS-70/MoS<sub>2</sub>. Insets of this figure are the TEM image (<b>left</b>) and histogram (<b>right</b>) of BGQS. The white circle indicates the particle size of B-GQDs. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B118-processes-12-01684" class="html-bibr">118</a>]. Copyright (2019), Frontiers.</p>
Full article ">Figure 12
<p>(<b>a</b>) Schematic illustration of ZAB. (<b>b</b>) Charging–discharging curves and (<b>c</b>) power density curves of alkaline and neutral ZAB with Fe-NBrGO. (<b>d</b>) Stability test and (<b>e</b>) voltaic efficiency of alkaline and neutral ZAB with Fe-NBrGO. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B114-processes-12-01684" class="html-bibr">114</a>]. Copyright (2023), MDPI.</p>
Full article ">Figure 13
<p>Morphology and microstructure of the Fe/Fe<sub>3</sub>C/G nanocomposite. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B119-processes-12-01684" class="html-bibr">119</a>]. Copyright (2022), ACS.</p>
Full article ">Figure 14
<p>TEM images of GO (<b>a</b>), GO/FePc (<b>b</b>), GO/MnPc (<b>c</b>), and GO/CoPc (<b>d</b>). Reproduced with permission from ref. [<a href="#B121-processes-12-01684" class="html-bibr">121</a>]. Copyright © 2024, Elsevier.</p>
Full article ">Figure 15
<p>(<b>a</b>) SEM morphology of 3D-N/S and (<b>b</b>) 3D-Fe/N/S. Reproduced with permission from ref. [<a href="#B122-processes-12-01684" class="html-bibr">122</a>]. Copyright © 2018, Elsevier.</p>
Full article ">Figure 16
<p>SEM and TEM images of Co/N/rGO(NH<sub>3</sub> ) (<b>a</b>,<b>b</b>) SEM images of Co/N/rGO(NH<sub>3</sub>) powder. Reproduced with permission from ref. [<a href="#B124-processes-12-01684" class="html-bibr">124</a>]. Copyright © 2013, Elsevier.</p>
Full article ">Figure 17
<p>(<b>a</b>,<b>b</b>) TEM images and (<b>c</b>) HR-TEM images of the Co/CoO-NGA. Reproduced under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B130-processes-12-01684" class="html-bibr">130</a>]. Copyright (2016), Frontiers.</p>
Full article ">Figure 18
<p>(<b>a</b>) TEM image of Co/Co<sub>3</sub>O<sub>4</sub>@NCs (<b>b</b>) EDS elemental mapping images of Co/Co<sub>3</sub>O<sub>4</sub>@NCs. Reproduced with permission from ref. [<a href="#B132-processes-12-01684" class="html-bibr">132</a>]. Copyright © 2020, Elsevier.</p>
Full article ">Figure 19
<p>SEM images of (<b>a</b>) NG and (<b>b</b>) FeCoOx@NG. Repeated under the terms of the CC-BY Creative Commons Attribution 4.0 International license [<a href="#B128-processes-12-01684" class="html-bibr">128</a>]. Copyright (2022), Wiley.</p>
Full article ">Figure 20
<p>TEM image of (<b>a</b>) CMO/rGO and (<b>b</b>) CMO/N-rGO. Reproduced with permission from ref. [<a href="#B133-processes-12-01684" class="html-bibr">133</a>]. Copyright © 2014, Elsevier.</p>
Full article ">Figure 21
<p>SEM images of (<b>a</b>) CoS<sub>2</sub>/N, S-GO, (<b>b</b>) CoS<sub>2</sub>, and (<b>c</b>) N, S-GO. Reproduced with permission from ref. [<a href="#B136-processes-12-01684" class="html-bibr">136</a>]. Copyright © 2015, Elsevier.</p>
Full article ">
12 pages, 5096 KiB  
Article
Theoretical Analysis of Superior Photodegradation of Methylene Blue by Cerium Oxide/Reduced Graphene Oxide vs. Graphene
by Nguyen Hoang Hao, Phung Thi Lan, Nguyen Ngoc Ha, Le Minh Cam and Nguyen Thi Thu Ha
Molecules 2024, 29(16), 3821; https://doi.org/10.3390/molecules29163821 - 12 Aug 2024
Viewed by 201
Abstract
Density functional theory and a semi-empirical quantum chemical approach were used to evaluate the photocatalytic efficiency of ceria (CeO2) combined with reduced graphene oxide (rGO) and graphene (GP) for degrading methylene blue (MB). Two main aspects were examined: the adsorption ability [...] Read more.
Density functional theory and a semi-empirical quantum chemical approach were used to evaluate the photocatalytic efficiency of ceria (CeO2) combined with reduced graphene oxide (rGO) and graphene (GP) for degrading methylene blue (MB). Two main aspects were examined: the adsorption ability of rGO and GP for MB, and the separation of photogenerated electrons and holes in CeO2/rGO and CeO2/GP. Our results, based on calculations of the adsorption energy, population analysis, bond strength index, and reduced density gradient, show favorable energetics for MB adsorption on both rGO and GP surfaces. The process is driven by weak, non-covalent interactions, with rGO showing better MB adsorption. A detailed analysis involving parameters like fractional occupation density, the centroid distance between molecular orbitals, and the Lewis acid index of the catalysts highlights the effective charge separation in CeO2/rGO compared to CeO2/GP. These findings are crucial for understanding photocatalytic degradation mechanisms of organic dyes and developing efficient photocatalysts. Full article
(This article belongs to the Topic Advances in Computational Materials Sciences)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Models of GP (<b>a</b>) and rGO (<b>b</b>); color codes: brown—C; ivory—H; red—O.</p>
Full article ">Figure 2
<p>Optimized adsorption configurations of MB on GP (<b>a</b>) and rGO (<b>b</b>); color codes: brown—C; ivory—H, yellow—Ce; red—O; green—Cl; gray—N; light yellow—S.</p>
Full article ">Figure 3
<p>IBSI values corresponding to the interatomic interactions between atoms of MB and atoms of rGO.</p>
Full article ">Figure 4
<p>Scatter graph of RDG for MB/GP (<b>a</b>) and MB/rGO (<b>b</b>) systems.</p>
Full article ">Figure 5
<p>RDG isosurfaces (isovalue = 0.8) of MB/GP (<b>a</b>) and MB/rGO (<b>b</b>) systems.</p>
Full article ">Figure 6
<p>Optimized structures of CeO<sub>2</sub>/GP (<b>a</b>) and CeO<sub>2</sub>/rGO (<b>b</b>).</p>
Full article ">Figure 7
<p>Frontier molecular orbitals of CeO<sub>2</sub>/GP: HOMO (<b>a</b>) and LUMO (<b>b</b>); and CeO<sub>2</sub>/rGO: HOMO (<b>c</b>) and LUMO (<b>d</b>).</p>
Full article ">Figure 8
<p>The FOD maps of GP, rGO, CeO<sub>2</sub>/GP, and CeO<sub>2</sub>/rGO.</p>
Full article ">
15 pages, 5440 KiB  
Article
Enhancing Flame Retardancy and Smoke Suppression in EPDM Rubber Using Sepiolite-Based Systems
by Jiawang Zheng, Xu Zhang, Dawei Liu, Liwei Zhang, Yuxia Guo, Wei Liu, Shuai Zhao and Lin Li
Polymers 2024, 16(16), 2281; https://doi.org/10.3390/polym16162281 - 12 Aug 2024
Viewed by 273
Abstract
The burning of Ethylene–Propylene–Diene Monomer (EPDM) rubber generates substantial smoke, posing a severe threat to the environment and personal safety. Considering the growing emphasis on safety and environmental protection, conventional non-smoke-suppressing flame retardants no longer satisfy the present application requirements. Consequently, there is [...] Read more.
The burning of Ethylene–Propylene–Diene Monomer (EPDM) rubber generates substantial smoke, posing a severe threat to the environment and personal safety. Considering the growing emphasis on safety and environmental protection, conventional non-smoke-suppressing flame retardants no longer satisfy the present application requirements. Consequently, there is an urgent need to develop a novel flame retardant capable of suppressing smoke formation while providing flame retardancy. Sepiolite (SEP), a porous silicate clay mineral abundant in silica and magnesium, exhibits notable advantages in the realm of flame retardancy and smoke suppression. This research focuses on the synthesis of two highly efficient flame-retardant smoke suppression systems, namely AEGS and PEGS, using Enteromorpha (EN), graphene (GE), sepiolite (SEP), ammonium polyphosphate (APP), and/or piperazine pyrophosphate (PPAP). The studied flame-retardant systems were then applied to EPDM rubber and the flame-retardant and smoke suppression abilities of EPDM/AEGS and EPDM/PEGS composites were compared. The findings indicate that the porous structure of sepiolite plays a significant role in reducing smoke emissions for EPDM composites during combustion. Full article
Show Figures

Figure 1

Figure 1
<p>Mode of Action for AEGS and PEGS in EPDM.</p>
Full article ">Figure 2
<p>FTIR spectra of SEP and K-SEP.</p>
Full article ">Figure 3
<p>AEGS3 (<b>a</b>) and PEGS3 (<b>b</b>) vertical combustion process photos.</p>
Full article ">Figure 4
<p>HRR (<b>a</b>), THR <b>(b</b>), SPR (<b>c</b>), and TSP (<b>d</b>) of EPDM/AEGS and PEGS composite materials.</p>
Full article ">Figure 5
<p>CO production rate (<b>a</b>), CO<sub>2</sub> production rate (<b>b</b>) and Mass (<b>c</b>) of AEGS and PEGS composite materials.after CCT test.</p>
Full article ">Figure 6
<p>Digital photos of carbon residues of composite materials tested by cone calorimetry: (<b>a</b>) PEGS1, (<b>b</b>) PEGS2, (<b>c</b>) PEGS3, (<b>d</b>) AEGS1, (<b>e</b>) AEGS2, and (<b>f</b>) AEGS3.</p>
Full article ">Figure 7
<p>TG curve (<b>a</b>) and DTG curve (<b>b</b>) of EPDM/AEGS and EPDM/PEGS composites under the N<sub>2</sub> atmosphere.</p>
Full article ">Figure 8
<p>Raman spectra after band fitting of various EPDM composite carbon residues: (<b>a</b>) PEGS0, (<b>b</b>) PEGS1, (<b>c</b>) PEGS2, (<b>d</b>) PEGS3, (<b>e</b>) AEGS0, (<b>f</b>) AEGS1, (<b>g</b>) AEGS2, and (<b>h</b>) AEGS3.</p>
Full article ">
18 pages, 7451 KiB  
Article
Study on the Lubrication Performance of Graphene-Based Polyphosphate Lubricants in High-Temperature Steel–Steel Friction Pair
by Kaifu Mi, Qingqing Ding, Xiangru Xu, Yu Lei, Juncheng Wang and Ning Kong
Surfaces 2024, 7(3), 571-588; https://doi.org/10.3390/surfaces7030039 (registering DOI) - 11 Aug 2024
Viewed by 329
Abstract
In the study, a hybrid lubricant was prepared by introducing graphene into a polyphosphate lubricant. In the tribological test of a steel/steel friction pair at the high temperature of 800 °C, the addition of a small proportion of graphene significantly enhances the lubrication [...] Read more.
In the study, a hybrid lubricant was prepared by introducing graphene into a polyphosphate lubricant. In the tribological test of a steel/steel friction pair at the high temperature of 800 °C, the addition of a small proportion of graphene significantly enhances the lubrication performance of polyphosphate at elevated temperatures. The coefficient of friction and the wear were obviously held down while the surface quality of the high-temperature friction pair was enhanced effectively with the graphene-strengthened polyphosphate lubricant, compared with the dry sliding condition. Through scanning electron microscopy and Raman spectroscopy analysis, the formation mechanism of tribofilm and the antiwear performance of the hybrid lubricant are further explained. This lubricant effectively combines the advantages of both; the combination of polyphosphate melted at elevated temperature with graphene and metal surfaces ensures the self-sealing of the friction contact area and brings better high-temperature oxidation resistance. At the same time, the presence of graphene provides excellent strength to the friction film and ensures the anti-wear and wear-resistant performance of the lubricant at high temperatures. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Three formulated lubricants and their main components.</p>
Full article ">Figure 2
<p>The temperature regime and the conditions of the tribotests.</p>
Full article ">Figure 3
<p>The TG curves and SEM images of (<b>a</b>) G showing its thermal stability in air and N2 and micro-morphology of initial film; (<b>b</b>) PO showing its thermal stability in air and N2 and micro-morphology of initial film; (<b>c</b>) G-PO showing its thermal stability in air and N2 and micro-morphology of initial film.</p>
Full article ">Figure 4
<p>Friction coefficient vs. time curves of dry friction and tribotests under different lubrication conditions.</p>
Full article ">Figure 5
<p>Laser scanning confocal microscopy, 3D topography, and height profile of the grinding mark for (<b>a</b>) dry friction; (<b>b</b>) tribotest under G lubricant; (<b>c</b>) tribotest under PO lubricant; (<b>d</b>) tribotest under G-PO lubricant.</p>
Full article ">Figure 6
<p>Wear loss of steel/steel pair under unlubricated and different lubricated conditions at 800 °C.</p>
Full article ">Figure 7
<p>SEM images of wear scars and tracks on the ball and flats after tribotests without lubricants (<b>a</b>,<b>b</b>), with G (<b>c</b>,<b>d</b>), with PO (<b>e</b>,<b>f</b>), and with G-PO (<b>g</b>,<b>h</b>); the Raman spectra of points within the grinding tracks on the flats after tribotests with G (<b>d’</b>) and G-PO (<b>h’</b>).</p>
Full article ">Figure 8
<p>Iron oxides in the grinding marks on the plates formed during tests at 800 °C without lubricants (<b>a</b>); with G (<b>b</b>); with PO (<b>c</b>); and with G-PO (<b>d</b>).</p>
Full article ">Figure 9
<p>The gradually enlarged SEM images of the interior of the grinding mark after tribotest with G-PO: (<b>a</b>) SEM image of the grinding mark; (<b>b</b>) the detail image of (<b>a</b>); (<b>c</b>) the detail image of the red position in (<b>b</b>); (<b>e</b>) the detail image of the blue position in (<b>b</b>); (<b>d</b>) The further amplification image of (<b>c</b>) and (<b>f</b>) the further amplification image of (<b>e</b>).</p>
Full article ">Figure 10
<p>The gradually enlarged SEM images of the edge of the grinding mark after tribotest with G-PO: (<b>a</b>–<b>c</b>).</p>
Full article ">Figure 11
<p>EDS mappings and the corresponding EDS spectra of the particle in <a href="#surfaces-07-00039-f009" class="html-fig">Figure 9</a>f (<b>a</b>,<b>b</b>) and <a href="#surfaces-07-00039-f010" class="html-fig">Figure 10</a>c (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 12
<p>Lubrication mechanism of G-PO lubricant at various temperatures.</p>
Full article ">
28 pages, 3497 KiB  
Review
Polymer-Assisted Graphite Exfoliation: Advancing Nanostructure Preparation and Multifunctional Composites
by Jaime Orellana, Esteban Araya-Hermosilla, Andrea Pucci and Rodrigo Araya-Hermosilla
Polymers 2024, 16(16), 2273; https://doi.org/10.3390/polym16162273 - 10 Aug 2024
Viewed by 478
Abstract
Exfoliated graphite (ExG) embedded in a polymeric matrix represents an accessible, cost-effective, and sustainable method for generating nanosized graphite-based polymer composites with multifunctional properties. This review article analyzes diverse methods currently used to exfoliate graphite into graphite nanoplatelets, few-layer graphene, and polymer-assisted graphene. [...] Read more.
Exfoliated graphite (ExG) embedded in a polymeric matrix represents an accessible, cost-effective, and sustainable method for generating nanosized graphite-based polymer composites with multifunctional properties. This review article analyzes diverse methods currently used to exfoliate graphite into graphite nanoplatelets, few-layer graphene, and polymer-assisted graphene. It also explores engineered methods for small-scale pilot production of polymer nanocomposites. It highlights the chemistry involved during the graphite intercalation and exfoliation process, particularly emphasizing the interfacial interactions related to steric repulsion forces, van der Waals forces, hydrogen bonds, π-π stacking, and covalent bonds. These interactions promote the dispersion and stabilization of the graphite derivative structures in polymeric matrices. Finally, it compares the enhanced properties of nanocomposites, such as increased thermal and electrical conductivity and electromagnetic interference (EMI) shielding applications, with those of neat polymer materials. Full article
(This article belongs to the Special Issue Functional Graphene-Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>Graphite structures and sources.</p>
Full article ">Figure 2
<p>SEM micrographs of rolling intercalation PS with 5% of colloidal graphite [<a href="#B27-polymers-16-02273" class="html-bibr">27</a>]. Reproduced with permission from Tu, H.; Polymers for advanced technologies; Published by John/Wiley &amp; Sons Ltd.; 2008.</p>
Full article ">Figure 3
<p>Orientation of the graphite platelet structures in extruded strands: (left) schematic figure of platelet orientation along strand flow direction by extrusion out of the die; and (right) transmission light microscopy pictures of samples cut perpendicular to the strand direction (shows mainly the layer thickness) and cut parallel to the long-axis of the strand (shows the lateral dimension of visible GNP structures), here shown for 1 wt% Graphene nanopowder AO-3 in PC [<a href="#B124-polymers-16-02273" class="html-bibr">124</a>]. Reproduced with permission from Pötschke, P.; Materials; Published by MDPI; 2017.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>) PVDF/expanded graphite composite fabricated without water (P-EG) and (<b>b</b>) PVDF/expanded graphite composite fabricated with water (P-EG-W) samples [<a href="#B91-polymers-16-02273" class="html-bibr">91</a>]. Reproduced with permission from Tong, J.; Macromolecular materials and engineering; Published by Wiley-VCH Verlag GMBH &amp; Co.; 2020.</p>
Full article ">Figure 5
<p>FESEM images of (<b>A</b>) PCGF-30 [<a href="#B87-polymers-16-02273" class="html-bibr">87</a>]. Reproduced with permission from Pradhan, S.S.; Polymer composites; Published by John/Wiley &amp; Sons Ltd.; 2021.</p>
Full article ">Figure 6
<p>SEM images of graphite at magnifications of 15 and 150 Kx at different ball milling times: (<b>a</b>,<b>b</b>) 0 h; (<b>c</b>,<b>d</b>) 1 h; (<b>e</b>,<b>f</b>) 4 h; (<b>g</b>,<b>h</b>) 8 h; (<b>i</b>,<b>j</b>) 16 h [<a href="#B130-polymers-16-02273" class="html-bibr">130</a>]. Reproduced with permission from Visco, A.; Polymers; Published by MDPI; 2021.</p>
Full article ">Figure 7
<p>EMI-S effectiveness (dB) and wt% filler in different matrix.</p>
Full article ">Figure 8
<p>Thermal conductivity and wt% filler in different matrix, conductive nanocomposite.</p>
Full article ">Figure 9
<p>Thermal conductivity and wt% filler in different matrices, conductive nanocomposite.</p>
Full article ">Scheme 1
<p>Engineering state-of-the-art techniques to generate exfoliated graphite/polymers composites.</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 276
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 ">
10 pages, 4355 KiB  
Article
An Electrochemical Biosensor for the Detection of Pulmonary Embolism and Myocardial Infarction
by Yaw-Jen Chang, Fu-Yuan Siao and En-Yu Lin
Biosensors 2024, 14(8), 386; https://doi.org/10.3390/bios14080386 - 9 Aug 2024
Viewed by 285
Abstract
Due to the clinical similarities between pulmonary embolism (PE) and myocardial infarction (MI), physicians often encounter challenges in promptly distinguishing between them, potentially missing the critical window for the correct emergency response. This paper presents a biosensor, termed the PEMI biosensor, which is [...] Read more.
Due to the clinical similarities between pulmonary embolism (PE) and myocardial infarction (MI), physicians often encounter challenges in promptly distinguishing between them, potentially missing the critical window for the correct emergency response. This paper presents a biosensor, termed the PEMI biosensor, which is designed for the identification and quantitative detection of pulmonary embolism or myocardial infarction. The surface of the working electrode of the PEMI biosensor was modified with graphene oxide and silk fibroin to immobilize the mixture of antibodies. Linear sweep voltammetry was employed to measure the current-to-potential mapping of analytes, with the calculated curvature serving as a judgment index. Experimental results showed that the curvature exhibited a linear correlation with the concentration of antigen FVIII, and a linear inverse correlation with the concentration of antigen cTnI. Given that FVIII and cTnI coexist in humans, the upper and lower limits were determined from the curvatures of a set of normal concentrations of FVIII and cTnI. An analyte with a curvature exceeding the upper limit can be identified as pulmonary embolism, while a curvature falling below the lower limit indicates myocardial infarction. Additionally, the further the curvature deviates from the upper or lower limits, the more severe the condition. The PEMI biosensor can serve as an effective detection platform for physicians. Full article
(This article belongs to the Section Biosensors and Healthcare)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The design and dimensions of the biosensor. (<b>b</b>) The surface modification process. (<b>c</b>) The biofunctionalization process. Different colors represent different antibodies. (<b>d</b>) The electrical measurement process.</p>
Full article ">Figure 2
<p>The SEM images after surface modification and biofunctionalization process: (<b>a</b>) The modified GO surface in the form of a uniformly distributed mesh. (<b>b</b>) SF coating in a semi-translucent and flat form. (<b>c</b>) Antibodies adsorbed as fine spheroids.</p>
Full article ">Figure 3
<p>The measured current–voltage (I–V) curves using a high-accuracy electric meter (Keithley 2614B). The curvature of the fitted (red) curve is used as a judgment index.</p>
Full article ">Figure 4
<p>Detection performance by single antibody: (<b>a</b>) The curvature <math display="inline"><semantics> <mrow> <mi>A</mi> </mrow> </semantics></math> was linearly related to the single antibody concentration of anti-FVIII and anti-cTnI, respectively. (<b>b</b>) Electrical measurements of the layer-by-layer deposits on PEMI biosensor.</p>
Full article ">Figure 5
<p>PEMI biosensor with the mixture of antibodies: (<b>a</b>) Detection performance. (<b>b</b>) Maximum measurable concentration.</p>
Full article ">Figure 6
<p>(<b>a</b>) The curvatures of coexisting FVIII and cTnI: Inability to distinguish between normal and abnormal conditions. (<b>b</b>) Histogram with upper and lower limits to detect PE or MI.</p>
Full article ">
14 pages, 6351 KiB  
Article
Comparative Study on the Lubrication of Ti3C2TX MXene and Graphene Oxide Nanofluids for Titanium Alloys
by Yaru Tian, Ye Yang, Heyi Zhao, Lina Si, Hongjuan Yan, Zhaoliang Dou, Fengbin Liu and Yanan Meng
Lubricants 2024, 12(8), 285; https://doi.org/10.3390/lubricants12080285 - 9 Aug 2024
Viewed by 258
Abstract
Titanium alloys are difficult to machine and have poor tribological properties. Nanoparticles have good cooling and lubricating properties, which can be used in metal cutting fluid. The lubrication characteristics of the two-dimensional materials Ti3C2TX MXene and graphene oxide [...] Read more.
Titanium alloys are difficult to machine and have poor tribological properties. Nanoparticles have good cooling and lubricating properties, which can be used in metal cutting fluid. The lubrication characteristics of the two-dimensional materials Ti3C2TX MXene and graphene oxide in water-based fluid for titanium alloys were comparatively investigated in this paper. Graphene oxide had smaller friction coefficients and wear volume than Ti3C2TX MXene nanofluid. As to the mechanism, MXene easily formed TiO2 for the tribo-chemical reaction, which accelerated wear. Moreover, GO nanofluid can form a more uniform and stable friction layer between the frictional interface, which reduces the friction coefficient and decreases the adhesive wear. The effects of different surfactants on the lubricating properties of MXene were further investigated. It was found that the cationic surfactant Hexadecyl trimethyl ammonium chloride (1631) had the lowest friction coefficient and anti-wear properties for the strong electrostatic attraction with MXene nanoparticles. The results of this study indicate that 2D nanoparticles, especially graphene oxide, could improve the lubricating properties of titanium alloys. It provides insight into the application of water-based nanofluids for difficult-to-machine materials to enhance surface quality and cutting efficiency. The developed nanofluid, which can lubricate titanium alloys, effectively has very broad applications in prospect. Full article
(This article belongs to the Special Issue Advanced Polymeric and Colloidal Lubricants)
Show Figures

Figure 1

Figure 1
<p>The microtopography and EDS analysis of the nanoparticles (<b>a</b>) graphene oxide and (<b>b</b>) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>X</sub></span> MXene.</p>
Full article ">Figure 2
<p>(<b>a</b>) The schematic diagram and (<b>b</b>) the photo of the CFT-I tribo-tester.</p>
Full article ">Figure 3
<p>(<b>a</b>) Friction coefficient and (<b>b</b>) wear volume and average friction coefficients of the two kinds of nanofluids and CSS solutions.</p>
Full article ">Figure 4
<p>SEM morphology of the tracks on Ti-6Al-4V lubricated by (<b>a</b>) 10%CSS solution, (<b>b</b>) graphene oxide, and (<b>c</b>) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>X</sub></span> MXene.</p>
Full article ">Figure 5
<p>Friction coefficient of (<b>a</b>) GO and (<b>b</b>) Ti<sub>3</sub>C<sub>2</sub>T<span class="html-italic"><sub>X</sub></span> MXene nanofluids with different concentrations.</p>
Full article ">Figure 6
<p>XPS spectrograms C1s, Ti2p, and O1s after lubrication with (<b>a</b>–<b>c</b>) Ti<sub>3</sub>C<sub>2</sub>T<sub>X</sub> MXene and (<b>d</b>–<b>f</b>) GO nanofluids.</p>
Full article ">Figure 7
<p>(<b>a</b>) Friction coefficient and (<b>b</b>) wear volume and average friction coefficients of nanofluids with different surfactants.</p>
Full article ">Figure 8
<p>SEM morphology of the trajectories on Ti-6Al-4V, consisting of (<b>a</b>) Alkylphenol ethoxylates (APE-10), (<b>b</b>) Octaphenyl polyoxyethylene (OP-10), (<b>c</b>) Sodium dodecyl benzene sulfonate (SDBS), (<b>d</b>) Sodium dodecyl sulfate (SDS), and (<b>e</b>) Hexadecyl trimethyl ammonium chloride (1631).</p>
Full article ">
25 pages, 5560 KiB  
Review
Hydrogen Storage Properties of Metal-Modified Graphene Materials
by Leela Sotsky, Angeline Castillo, Hugo Ramos, Eric Mitchko, Joshua Heuvel-Horwitz, Brian Bick, Devinder Mahajan and Stanislaus S. Wong
Energies 2024, 17(16), 3944; https://doi.org/10.3390/en17163944 - 9 Aug 2024
Viewed by 243
Abstract
The absence of adequate methods for hydrogen storage has prevented the implementation of hydrogen as a major source of energy. Graphene-based materials have been considered for use as solid hydrogen storage, because of graphene’s high specific surface area. However, these materials alone do [...] Read more.
The absence of adequate methods for hydrogen storage has prevented the implementation of hydrogen as a major source of energy. Graphene-based materials have been considered for use as solid hydrogen storage, because of graphene’s high specific surface area. However, these materials alone do not meet the hydrogen storage standard of 6.5 wt.% set by the United States Department of Energy (DOE). They can, however, be easily modified through either decoration or doping to alter their chemical properties and increase their hydrogen storage capacity. This review is a compilation of various published reports on this topic and summarizes results from theoretical and experimental studies that explore the hydrogen storage properties of metal-modified graphene materials. The efficacy of alkali, alkaline earth metal, and transition metal decoration is examined. In addition, metal doping to further increase storage capacity is considered. Methods for hydrogen storage capacity measurements are later explained and the properties of an effective hydrogen storage material are summarized. Full article
(This article belongs to the Section C: Energy Economics and Policy)
Show Figures

Figure 1

Figure 1
<p>Structure of various carbon nanostructures including diamond, graphite, C<sub>60</sub>, CNTs, graphene, and 3D graphene–CNT hybrid materials [<a href="#B16-energies-17-03944" class="html-bibr">16</a>].</p>
Full article ">Figure 2
<p>The structure of graphene and its various derivatives: (<b>a</b>) pristine graphene lattice; (<b>b</b>) graphene oxide; and (<b>c</b>) reduced graphene oxide [<a href="#B19-energies-17-03944" class="html-bibr">19</a>]. Reproduced with permission.</p>
Full article ">Figure 3
<p>From left to right, Stone–Wales defect, single-vacancy defect, and double-vacancy defect within a graphene lattice, with each defect highlighted in red [<a href="#B21-energies-17-03944" class="html-bibr">21</a>]. Reproduced with permission.</p>
Full article ">Figure 4
<p>Optimized structures of H<sub>2</sub> adsorbed onto Li-decorated graphene. (<b>a</b>) Li atom decorated at the center of the hexagonal pore of graphene. (<b>b</b>–<b>f</b>) Graphene in 3 × 3 formation with H<sub>2</sub> molecules adsorbed. Li atoms are represented by purple-colored molecules;, hydrogen atoms are denoted by white-colored molecules; and carbon atoms of graphene are highlighted by gray-colored molecules [<a href="#B33-energies-17-03944" class="html-bibr">33</a>]. Reproduced with permission.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>f</b>) Optimized structures of 1–6 hydrogen molecules adsorbed onto single-sided Na-decorated boron-substituted graphene (BSG). The outer- and inner-layer hydrogen molecules are represented by yellow and white molecules, respectively. Sodium, boron, and carbon atoms are represented by purple, pink, and gray molecules, respectively [<a href="#B35-energies-17-03944" class="html-bibr">35</a>]. Reproduced with permission.</p>
Full article ">Figure 6
<p>Molecular model of Sc decorated on N-doped graphene with four H<sub>2</sub> adsorbed<sub>,</sub> created using the Avogadro Molecular Software: an open-source molecular builder and visualization tool, used as version 1.2.0. Carbon, nitrogen, and hydrogen atoms are represented by dark gray, blue, and light-gray molecules, respectively [<a href="#B46-energies-17-03944" class="html-bibr">46</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) STM image of 0.55 ML (monolayer) titanium nanoparticles dispersed onto a pristine graphene surface. (<b>b</b>) STM image of titanium nanoparticles after sputtering the sample at E  =  300 eV for 150 s [<a href="#B49-energies-17-03944" class="html-bibr">49</a>]. Reproduced with permission.</p>
Full article ">Figure 8
<p>Schematic of Ru-decorated SWCNT + <span class="html-italic">n</span>H<sub>2</sub> systems (where <span class="html-italic">n</span> = 1–4). (<b>A</b>) Ru/SWCNT + H<sub>2</sub>, (<b>B</b>) Ru/SWCNT + 2H<sub>2</sub>, (<b>C</b>) Ru/SWCNT + 3H<sub>2</sub>, and (<b>D</b>) Ru/SWCNT + 4H<sub>2</sub>. The upper images depict the optimized structures. Carbon, ruthenium, and hydrogen atoms are represented by green, blue, and red molecules, respectively. The lower images represent the isosurfaces of optimized systems (isosurface level = 0.09875 e Å<sup>−3</sup>). Carbon, ruthenium, and hydrogen atoms are represented by green, gray, and orange molecules, respectively [<a href="#B61-energies-17-03944" class="html-bibr">61</a>]. Reproduced with permission.</p>
Full article ">Figure 9
<p>Hydrogen capacities of modified graphene materials at 303 K [<a href="#B63-energies-17-03944" class="html-bibr">63</a>]. Reproduced with permission.</p>
Full article ">Figure 10
<p>TEM images of (<b>a</b>) Pt/RGO and (<b>b</b>) Pt/NRGO. (<b>c</b>,<b>d</b>) The size distributions of Pt nanoparticles on Pt/RGO and Pt/NRGO, respectively, are shown [<a href="#B66-energies-17-03944" class="html-bibr">66</a>]. Reproduced with permission.</p>
Full article ">Figure 11
<p>Schematic diagram of a typical Sieverts apparatus, wherein V<sub>cell</sub> is the volume of the cell containing the sample material and V<sub>ref</sub> is the volume of the reference cell [<a href="#B71-energies-17-03944" class="html-bibr">71</a>]. Reproduced with permission.</p>
Full article ">
17 pages, 4274 KiB  
Article
ZnO–Graphene Oxide Nanocomposite for Paclitaxel Delivery and Enhanced Toxicity in Breast Cancer Cells
by Lorenzo Francesco Madeo, Christine Schirmer, Giuseppe Cirillo, Ayah Nader Asha, Rasha Ghunaim, Samuel Froeschke, Daniel Wolf, Manuela Curcio, Paola Tucci, Francesca Iemma, Bernd Büchner, Silke Hampel and Michael Mertig
Molecules 2024, 29(16), 3770; https://doi.org/10.3390/molecules29163770 - 9 Aug 2024
Viewed by 383
Abstract
A ZnO-Graphene oxide nanocomposite (Z-G) was prepared in order to exploit the biomedical features of each component in a single anticancer material. This was achieved by means of an environmentally friendly synthesis, taking place at a low temperature and without the involvement of [...] Read more.
A ZnO-Graphene oxide nanocomposite (Z-G) was prepared in order to exploit the biomedical features of each component in a single anticancer material. This was achieved by means of an environmentally friendly synthesis, taking place at a low temperature and without the involvement of toxic reagents. The product was physicochemically characterized. The ZnO-to-GO ratio was determined through thermogravimetric analysis, while scanning electron microscopy and transmission electron microscopy were used to provide insight into the morphology of the nanocomposite. Using energy-dispersive X-ray spectroscopy, it was possible to confirm that the graphene flakes were homogeneously coated with ZnO. The crystallite size of the ZnO nanoparticles in the new composite was determined using X-ray powder diffraction. The capacity of Z-G to enhance the toxicity of the anticancer drug Paclitaxel towards breast cancer cells was assessed via a cell viability study, showing the remarkable anticancer activity of the obtained system. Such results support the potential use of Z-G as an anticancer agent in combination with a common chemotherapeutic like Paclitaxel, leading to new chemotherapeutic formulations. Full article
(This article belongs to the Special Issue Carbon Materials in Materials Chemistry—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>TGA of GO, ZnO NPs, Z-G, and Z-G*.</p>
Full article ">Figure 2
<p>FTIR spectra of GO, ZnO NPs, Z-G and Z-G*.</p>
Full article ">Figure 3
<p>pXRD pattern of ZnO NPs, Z-G, and Z-G*.</p>
Full article ">Figure 4
<p>SEM (<b>a</b>,<b>b</b>) and TEM (<b>c</b>–<b>e</b>) images of Z-G* (<b>a</b>) and Z-G (<b>b</b>–<b>e</b>). (<b>c</b>) Bright-field TEM (BFTEM) image of ZnO NPs clusters on GO sheets. The wrinkling of the latter is visible as dark lines. (<b>d</b>) BFTEM images at higher magnification showing morphology, size, and arrangement of the ZnO NPs within the cluster. (<b>e</b>) High-resolution TEM (HRTEM) images of ZnO NPs at the edge of the cluster showing lattice planes at a few NPs. (<b>f</b>) Fourier transform of (<b>e</b>) revealing reflections of both ZnO and GO.</p>
Full article ">Figure 5
<p>EDX mapping of Z-G* (<b>a</b>–<b>d</b>) and Z-G (<b>e</b>–<b>h</b>). The original SEM images are shown in (<b>a</b>,<b>e</b>). The detected elements were O (<b>b</b>,<b>f</b>), C (<b>c</b>,<b>g</b>), and Zn (<b>d</b>,<b>h</b>). Scale bar = 1 μm.</p>
Full article ">Figure 6
<p>Ptx release profiles for Z-G, GO, and ZnO NP samples.</p>
Full article ">Figure 7
<p>Cancer cells’ viability after 48 h of incubation with unloaded NPs (GO, Z-G), Ptx, and loaded NPs (ZnO/Ptx, GO/Ptx, and Z-G/Ptx) compared to the control (DMSO treatment). ***/**/* decreased viability vs. control (DMSO); ###/##/# decreased viability vs unloaded NPs; +++/++ decreased viability vs Ptx; */# <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.</p>
Full article ">Figure 8
<p>Light microscopy images of cancer cells treated with Ptx (50 μg/mL) and Z-G/Ptx (25 μg/mL) for 48 h. Scale bar = 100 µm.</p>
Full article ">
17 pages, 3234 KiB  
Article
Graphene Oxide Covalently Functionalized with 5-Methyl-1,3,4-thiadiazol-2-amine for pH-Sensitive Ga3+ Recovery in Aqueous Solutions
by Xi Zhu, Yong Guo and Baozhan Zheng
Molecules 2024, 29(16), 3768; https://doi.org/10.3390/molecules29163768 - 9 Aug 2024
Viewed by 259
Abstract
A novel graphene-based composite, 5-methyl-1,3,4-thiadiazol-2-amine (MTA) covalently functionalized graphene oxide (GO-MTA), was rationally developed and used for the selective sorption of Ga3+ from aqueous solutions, showing a higher adsorption capacity (48.20 mg g−1) toward Ga3+ than In3+ (15.41 [...] Read more.
A novel graphene-based composite, 5-methyl-1,3,4-thiadiazol-2-amine (MTA) covalently functionalized graphene oxide (GO-MTA), was rationally developed and used for the selective sorption of Ga3+ from aqueous solutions, showing a higher adsorption capacity (48.20 mg g−1) toward Ga3+ than In3+ (15.41 mg g−1) and Sc3+ (~0 mg g−1). The adsorption experiment’s parameters, such as the contact time, temperature, initial Ga3+ concentration, solution pH, and desorption solvent, were investigated. Under optimized conditions, the GO-MTA composite displayed the highest adsorption capacity of 55.6 mg g−1 toward Ga3+. Moreover, a possible adsorption mechanism was proposed using various characterization methods, including scanning electron microscopy (SEM) equipped with X-ray energy-dispersive spectroscopy (EDS), elemental mapping analysis, Fourier transform infrared (FT-IR) spectroscopy, and X-ray photoelectron spectroscopy (XPS). Ga3+ adsorption with the GO-MTA composite could be better described by the linear pseudo-second-order kinetic model (R2 = 0.962), suggesting that the rate-limiting step may be chemical sorption or chemisorption through the sharing or exchange of electrons between the adsorbent and the adsorbate. Importantly, the calculated qe value (55.066 mg g−1) is closer to the experimental result (55.60 mg g−1). The well-fitted linear Langmuir isothermal model (R2 = 0.972~0.997) confirmed that an interfacial monolayer and cooperative adsorption occur on a heterogeneous surface. The results showed that the GO-MTA composite might be a potential adsorbent for the enrichment and/or separation of Ga3+ at low or ultra-low concentrations in aqueous solutions. Full article
(This article belongs to the Special Issue Design and Application Based on Versatile Nano-Composites)
Show Figures

Figure 1

Figure 1
<p>SEM images of the samples: (<b>A</b>) GO; (<b>B</b>) GO-MTA composite; and (<b>C</b>) GO-MTA composite post Ga<sup>3+</sup> adsorption.</p>
Full article ">Figure 2
<p>Elemental-mapping images of the GO-MTA composite post Ga<sup>3+</sup> adsorption—(<b>A</b>) area tested, (<b>B</b>) C, (<b>C</b>) N, (<b>D</b>) O, (<b>E</b>) S, and (<b>F</b>) Ga<sup>3+</sup>—and (<b>G</b>) EDS of the composite.</p>
Full article ">Figure 3
<p>FT-IR spectra of the samples: (<b>A</b>) MTA and (<b>B</b>) GO-MTA composite.</p>
Full article ">Figure 4
<p>XPS spectra of the GO-MTA composite pre and post Ga<sup>3+</sup> adsorption: (<b>A</b>) survey spectra and XPS peak-differentiation-imitating analyses of (<b>B</b>) C 1s, (<b>C</b>) N 1s, (<b>D</b>) O 1s, and (<b>E</b>) S 2p.</p>
Full article ">Figure 5
<p>Adsorption properties of the GO-MTA composite for Ga<sup>3+</sup>, Sc<sup>3+</sup>, and In<sup>3+</sup> (dosage of adsorbent = 5.0 mg, <span class="html-italic">V</span> = 20.0 mL, <span class="html-italic">C</span><sub>0</sub> = 50.0 mg L<sup>−1</sup>, <span class="html-italic">T</span> = 25 °C, and <span class="html-italic">t</span> = 180 min).</p>
Full article ">Figure 6
<p>The adsorption capacity in MTA: (<b>A</b>) effect of the contact time (adsorbent dosage = 5.0 mg, <span class="html-italic">C</span><sub>0</sub> = 50.0 mg L<sup>−1</sup>, <span class="html-italic">T</span> = 298 K, and <span class="html-italic">V</span> = 20.0 mL; RSD = 0.14%~3.45%); (<b>B</b>) fitted using the linear pseudo-first-order kinetic model; (<b>C</b>) fitted using the nonlinear pseudo-first-order kinetic model; (<b>D</b>) fitted using the linear pseudo-second-order kinetic model; and (<b>E</b>) fitted using the nonlinear pseudo-second-order kinetic model.</p>
Full article ">Figure 7
<p>(<b>A</b>) Effects of the pH on the Ga<sup>3+</sup> adsorption capacity of the GO-MTA composite (adsorbent dosage = 5.0 mg, <span class="html-italic">C</span><sub>0</sub> = 50.0 mg L<sup>−1</sup>, <span class="html-italic">T</span> = 298 K, <span class="html-italic">V</span> = 20.0 mL, and <span class="html-italic">t</span> = 120 min; RSD = 0.25%~3.04%). (<b>B</b>) Zeta potential of the GO-MTA composite in the pH range of 2.0–6.0.</p>
Full article ">Figure 8
<p>Effects of the contact time on the GO-MTA composite’s Ga<sup>3+</sup> adsorption capacity: (<b>a</b>) effect of the initial concentration and temperature on the adsorption capacity of the GO-MTA (adsorbent dosage = 5.0 mg, <span class="html-italic">C</span><sub>0</sub> = 50.0 mg L<sup>−1</sup>, <span class="html-italic">T</span> = 298 K, <span class="html-italic">V</span> = 20.0 mL, <span class="html-italic">t</span> = 120 min, and pH = 3.0; RSD = 0.32%~2.12%); (<b>b</b>) fitted curve of the linear Langmuir isothermal model; (<b>c</b>) fitted curve of the nonlinear Langmuir isothermal model; (<b>d</b>) fitted curve of the linear Freundlich isothermal model; and (<b>e</b>) fitted curve of the nonlinear Freundlich isothermal model.</p>
Full article ">Figure 9
<p>The reusability of the GO-MTA composite.</p>
Full article ">Scheme 1
<p>A schematic diagram of the functionalization of the GO-MTA composite and its Ga<sup>3+</sup> adsorption mechanism.</p>
Full article ">
Back to TopTop