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

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,792)

Search Parameters:
Keywords = copper ion

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 4934 KiB  
Article
Cu-Ion Hybrid Porous Carbon with Nanoarchitectonics Derived from Heavy-Metal-Contaminated Biomass as Ultrahigh-Performance Supercapacitor
by Jieni Wang, Xiaobo Han, Shuqin Zhang, Haodong Hou, Chenlin Wei, Chenxiao Liu, Leichang Cao, Jinglai Zhang, Li Wang and Shicheng Zhang
Int. J. Mol. Sci. 2025, 26(2), 569; https://doi.org/10.3390/ijms26020569 - 10 Jan 2025
Abstract
It is challenging to handle heavy-metal-rich plants that grow in contaminated soil. The role of heavy metals in biomass on the physicochemical structure and electrochemical properties of their derived carbon has not been considered in previous research. In this study, Cu-ion hybrid nanoporous [...] Read more.
It is challenging to handle heavy-metal-rich plants that grow in contaminated soil. The role of heavy metals in biomass on the physicochemical structure and electrochemical properties of their derived carbon has not been considered in previous research. In this study, Cu-ion hybrid nanoporous carbon (CHNC) is prepared from Cu content-contaminated biomass through subcritical hydrocharization (HTC) coupling pyrolytic activation processes. The CHNCs are used as advanced electrode material for energy storage applications, exhibiting an impressively ultrahigh capacitance of 562 F g−1 at a current density of 1 A g−1 (CHNC-700-4-25), excellent energy density of 26.15 W h kg−1, and only 7.59% capacitance loss after enduring 10,000 cycles at a current density of 10 A g−1, making CHNCs rank in the forefront of previously known carbon-based supercapacitor materials. These comprehensive characterizations demonstrate that copper ions introduce new electrochemically active sites and enhance the conductivity and charge transport performance of the electrode material, elevating the specific capacitance of CHNC from 463 to 562 F g−1. These findings offer valuable insights into the effective energy storage application of heavy-metal-contaminated biomass wastes. Full article
Show Figures

Figure 1

Figure 1
<p>Van Kraveren diagram of hydrochars and porous carbons.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) CHNC-700-4-0, (<b>b</b>) CHNC-700-4-25, (<b>c</b>) CHNC-700-4-50, (<b>d</b>) CHNC-600-4-25, (<b>e</b>) CHNC-800-4-25, (<b>f</b>) CHNC-700-2-25, and (<b>g</b>) PSPH-25. (<b>h</b>) The dark-field SEM mapping images of CHNC-700-4-25 with C, N, O, and Cu elements.</p>
Full article ">Figure 3
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isotherms of CHNC-600-4-25, CHNC-700-4-25, and CHNC-800-4-25; (<b>b</b>) N<sub>2</sub> adsorption–desorption isotherms of CHNC-700-2-25, CHNC-700-4-0, and CHNC-700-4-25, and CHNC-700-4-50; (<b>c</b>) pore size distribution of CHNC-600-4-25, CHNC-700-4-25, and CHNC-800-4-25; (<b>d</b>) pore size distribution of CHNC-700-2-25, CHNC-700-4-0, and CHNC-700-4-25, and CHNC-700-4-50.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD pattern of all samples, (<b>b</b>) Raman spectra of CHNC-700-2-25, CHNC-700-4-0, CHNC-700-4-25, and CHNC-700-4-50; (<b>c</b>) full-scan XPS spectra of CHNC-700-4-0 and CHNC-700-4-25, (<b>d</b>–<b>f</b>) are high-resolution XPS spectra of Cu 1s, C 1s, and O 1s for CHNC-700-4-25, respectively.</p>
Full article ">Figure 5
<p>Electrochemical performance of samples in the three-electrode system: (<b>a</b>,<b>b</b>) GCD curves for all samples, (<b>c</b>,<b>d</b>) CV curves for all samples, (<b>e</b>) GCD curves of CHNC-700-4-25 at different densities from 0.5 to 20 A g<sup>−1</sup>, and (<b>f</b>) CV curves of CHNC-700-4-25 at different scanning rates from 5 to 200 mV s<sup>−1</sup>.</p>
Full article ">Figure 6
<p>(<b>a</b>) The specific capacitance of all samples at different current densities: (<b>b</b>) The electrochemical cycle test of CHNC-700-4-25 was tested at 10 A g<sup>−1</sup>, (<b>c</b>) the equivalent circuit diagram of CHNC-700-4-25, and (<b>d</b>) the Nyquist plots of three samples of different heavy-metal contents.</p>
Full article ">Figure 7
<p>(<b>a</b>) CV curves of CHNC-700-4-25 at different open-circuit voltages, (<b>b</b>) CV curves of CHNC-700-4-25 at 1.3 V at open-circuit voltage, (<b>c</b>) GCD curves of CHNC-700-4-25 at different current densities, and (<b>d</b>) the relationship of CHNC-700-4-25 between energy density and power density.</p>
Full article ">Figure 8
<p>Schematic diagram of the preparation process of Cu-ion hybrid nanoporous carbon (CHNC).</p>
Full article ">
15 pages, 6277 KiB  
Article
Impact of Ag Coating Thickness on the Electrochemical Behavior of Super Duplex Stainless Steel SAF2507 for Enhanced Li-Ion Battery Cases
by Hyeongho Jo, Jung-Woo Ok, Yoon-Seok Lee, Sanghun Lee, Yonghun Je, Shinho Kim, Seongjun Kim, Jinyong Park, Jonggi Hong, Taekyu Lee, Byung-Hyun Shin, Jang-Hee Yoon and Yangdo Kim
Crystals 2025, 15(1), 62; https://doi.org/10.3390/cryst15010062 - 9 Jan 2025
Viewed by 200
Abstract
Li-ion batteries are at risk of explosions caused by fires, primarily because of the high energy density of Li ions, which raises the temperature. Battery cases are typically made of plastic, aluminum, or SAF30400. Although plastic and aluminum aid weight reduction, their strength [...] Read more.
Li-ion batteries are at risk of explosions caused by fires, primarily because of the high energy density of Li ions, which raises the temperature. Battery cases are typically made of plastic, aluminum, or SAF30400. Although plastic and aluminum aid weight reduction, their strength and melting points are low. SAF30400 offers excellent strength and corrosion resistance but suffers from work hardening and low high-temperature strength at 700 °C. Additionally, Ni used for plating has a low current density of 25% international copper alloy standard (ICAS). SAF2507 is suitable for use as a Li-ion battery case material because of its excellent strength and corrosion resistance. However, the heterogeneous microstructure of SAF2507 after casting and processing decreases the corrosion resistance, so it requires solution heat treatment. To address these issues, in this study, SAF2507 (780 MPa, 30%) is solution heat-treated at 1100 °C after casting and coated with Ag (ICAS 108.4%) using physical vapor deposition (PVD). Ag is applied at five different thicknesses: 0.5, 1.0, 1.5, 2.0, and 2.5 μm. The surface conditions and electrochemical properties are then examined for each coating thickness. The results indicate that the PVD-coated surface forms a uniform Ag layer, with electrical conductivity increasing from 1.9% ICAS to 72.3% ICAS depending on the Ag coating thickness. This enhancement in conductivity can improve Li-ion battery safety on charge and use. This result is expected to aid the development of advanced Li-ion battery systems in the future. Full article
(This article belongs to the Special Issue Advances in Surface Modifications of Metallic Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the preparation and analysis timeline of Ag-coated super duplex stainless steel SAF2507: (# α) casting for manufacturing (red arrow), (# β) solution annealing to achieve homogeneous grains (red arrow), (# γ) Ag coating applied via PVD in thicknesses ranging from 0.0 to 2.5 μm (blue arrow), and (# δ) analysis of electrochemical behavior (green arrow).</p>
Full article ">Figure 2
<p>FE-SEM images illustrating the manufacturing process of super duplex stainless steel SAF2507: (<b>a</b>) casting and (<b>b</b>) solution annealing at 1100 °C.</p>
Full article ">Figure 3
<p>Volume fractions of austenite and ferrite in super duplex stainless steel SAF2507 for various manufacturing processes.</p>
Full article ">Figure 4
<p>Surface images of Ag-coated super duplex stainless steel SAF2507 with varying Ag coating thicknesses for enhanced Li-ion battery case applications: (<b>a</b>) coating thickness = 0.0 μm (before coating), (<b>b</b>) coating thickness = 0.5 μm, (<b>c</b>) coating thickness = 1.0 μm, (<b>d</b>) coating thickness = 1.5 μm, (<b>e</b>) coating thickness = 2.0 μm, and (<b>f</b>) coating thickness = 2.5 μm.</p>
Full article ">Figure 5
<p>XRD patterns for SDSS SAF2507 with varying Ag coating thicknesses for enhanced Li-ion battery cases: (<b>a</b>) intensity from 0 to 250,000 and (<b>b</b>) intensity from 0 to 5000.</p>
Full article ">Figure 5 Cont.
<p>XRD patterns for SDSS SAF2507 with varying Ag coating thicknesses for enhanced Li-ion battery cases: (<b>a</b>) intensity from 0 to 250,000 and (<b>b</b>) intensity from 0 to 5000.</p>
Full article ">Figure 6
<p>Surface roughness of Ag-coated super duplex stainless steel SAF2507 at varying coating thicknesses from 0 to 2.5 μm: (<b>a</b>) Ra (μm) and (<b>b</b>) roughness gap, defined as the difference between the maximum and minimum roughness (μm).</p>
Full article ">Figure 7
<p>GDS results illustrating the relationship between thickness (μm) and the concentration of major alloying elements (%) in SDSS SAF2507 with various Ag coating thicknesses, employed in enhanced Li-ion battery cases: (<b>a</b>) coating thickness = 0.0 μm (before coating), (<b>b</b>) coating thickness = 0.5 μm, (<b>c</b>) coating thickness = 1.0 μm, (<b>d</b>) coating thickness = 1.5 μm, (<b>e</b>) coating thickness = 2.0 μm, and (<b>f</b>) coating thickness = 2.5 μm.</p>
Full article ">Figure 8
<p>Electrical conductivity as a function of Ag coating thickness on SDSS SAF2507.</p>
Full article ">Figure 9
<p>Time (s) vs potential (V) curve, i.e., OCP curve for various Ag coating thicknesses on super duplex stainless steel SAF2507 in NaCl electrolyte solution of 3.5 wt.%.</p>
Full article ">Figure 10
<p>Potentiodynamic polarization curves displaying the relationship between potential (V) and current density (A/cm<sup>2</sup>) for SDSS SAF2507 with varying Ag coating thicknesses.</p>
Full article ">Figure 11
<p>Image depicting chloride ion attack on Ag-coated SDSS SAF2507 in an electrolyte solution.</p>
Full article ">
29 pages, 5402 KiB  
Article
Neurotoxic Effect of Myricitrin in Copper-Induced Oxidative Stress Is Mediated by Increased Intracellular Ca2+ Levels and ROS/p53/p38 Axis
by Ignacija Vlašić, Antonio Krstačić-Galić, Anđela Horvat, Nada Oršolić, Anja Sadžak, Lucija Mandić, Suzana Šegota and Maja Jazvinšćak Jembrek
Antioxidants 2025, 14(1), 46; https://doi.org/10.3390/antiox14010046 - 3 Jan 2025
Viewed by 376
Abstract
Although commonly appreciated for their anti-oxidative and neuroprotective properties, flavonoids can also exhibit pro-oxidative activity, potentially reducing cell survival, particularly in the presence of metal ions. Disrupted copper homeostasis is a known contributor to neuronal dysfunction through oxidative stress induction. This study investigated [...] Read more.
Although commonly appreciated for their anti-oxidative and neuroprotective properties, flavonoids can also exhibit pro-oxidative activity, potentially reducing cell survival, particularly in the presence of metal ions. Disrupted copper homeostasis is a known contributor to neuronal dysfunction through oxidative stress induction. This study investigated the effects of myricitrin (1–20 μg/mL) on copper-induced toxicity (0.5 mM CuSO4) in the neuroblastoma SH-SY5Y cell line. At non-toxic concentrations, myricitrin exacerbated copper’s toxic effects. The myricitrin-induced decrease in survival was accompanied with increased reactive oxygen species (ROS) production, reduced superoxide dismutase activity, and a lower GSH/GSSG ratio. In combination with copper, myricitrin also activated caspase-3/7, promoted nuclear chromatin changes, and compromised membrane integrity. At the protein level, myricitrin upregulated p53 and PUMA expression. The toxic effects of myricitrin were alleviated by the p38 inhibitor SB203580, the intracellular calcium chelator BAPTA-AM, and the NMDA receptor blocker MK-801, highlighting the significant role of the ROS/p53/p38 axis in cell death and the critical involvement of calcium ions in apoptosis induction. The atomic force microscopy was used to assess the surface morphology and nanomechanical properties of SH-SY5Y cells, revealing changes following myricitrin treatment. This research highlights the toxic potential of myricitrin and emphasizes the need for caution when considering flavonoid supplementation in conditions with elevated copper levels. Full article
Show Figures

Figure 1

Figure 1
<p>Effect of myricitrin on the viability of SH-SY5Y cells under physiological conditions and following exposure to excess copper ions. SH-SY5Y cells were treated with myricitrin at concentrations of up to 20 µg/mL for 24 h. Cell viability was assessed using the MTT assay (<b>A</b>,<b>B</b>), crystal violet staining (<b>C</b>,<b>D</b>), and ATP content measurement (<b>E</b>). Morphological changes in SH-SY5Y cells were observed using the EVOS Floid Cell Imaging System (<b>F</b>) and were also photographed following crystal violet staining (<b>G</b>). Statistical analysis was performed using one-way ANOVA followed by Dunnett’s multiple comparison tests. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>c</sup> <span class="html-italic">p</span> &lt; 0.001, and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.0001 vs. 0 group (0.5 mM CuSO<sub>4</sub>). Data are presented as means ± standard deviation from 3 to 4 independent experiments (one-way ANOVA followed by post hoc Dunnett’s test).</p>
Full article ">Figure 2
<p>Effect of myricitrin on ROS production, SOD activity, and GSH/GSSG ratio. SH-SY5Y cells were treated with 1–20 µg/mL myricitrin alone or in the presence of 0.5 mM CuSO<sub>4</sub>. After 24 h, ROS levels were measured using 2′,7′-dichlorodihydrofluorescein diacetate (<b>A</b>,<b>B</b>). SOD activity was determined using a colorimetric assay. The rate of the reduction in superoxide anions in a reaction with molecular oxygen is related to xanthine oxidase activity that is inhibited by SOD (<b>C</b>). The GSH/GSSG ratio was determined using a luminescence-based system involving glutathione-S-transferase coupled to a luciferase reaction (<b>D</b>). Effect of N-acetylcysteine, a GSH precursor, on cell survival in SH-SY5Y treated with copper and 10 or 20 µg/mL myricitrin was determined by MTT assay (<b>E</b>). Results are presented as mean values ± standard deviation from 3 to 5 independent experiments performed in triplicate. <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.0001 compared to copper-only treatment (one-way ANOVA followed by post hoc Dunnett’s test for (<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>); Student’s <span class="html-italic">t</span>-test for (<b>C</b>)).</p>
Full article ">Figure 3
<p>Effects of copper and myricitrin co-treatment on caspase-3/7 activity, membrane integrity, and nuclear changes. SH-SY5Y cells were treated with 0.5 mM CuSO<sub>4</sub> and myricitrin (1–20 µg/mL), and caspase-3/7 activity was measured using a luminogenic substrate (<b>A</b>). Membrane damage was assessed by measuring LDH activity in the culture medium (<b>B</b>). Chromatin condensation was assessed using Hoechst 33342 staining, while propidium iodide staining was used to detect cells in late apoptosis/necrosis (<b>C</b>). Data are presented as mean values ± standard deviation from three independent experiments performed in duplicate (<b>A</b>,<b>C</b>) or triplicate (<b>B</b>). * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. control group; <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>c</sup> <span class="html-italic">p</span> &lt; 0.001, and <sup>d</sup> <span class="html-italic">p</span> &lt; 0.0001 vs. copper alone (one-way ANOVA followed by post hoc Dunnett’s test) and <sup>##</sup> <span class="html-italic">p</span> &lt; 0.01, <sup>###</sup> <span class="html-italic">p</span> &lt; 0.01 (Student’s <span class="html-italic">t</span>-test). Representative images obtained by staining with Hoechst 33342 (blue) and propidium iodide (red) were taken using the EVOS Floid Cell Imaging System (<b>D</b>).</p>
Full article ">Figure 4
<p>Effects of copper and myricitrin on the expression of apoptotic and oxidative stress-related proteins. Protein expression levels of p53 (<b>A</b>), PUMA (<b>B</b>), Bax (<b>C</b>), Bcl-2 (<b>D</b>), PARP-1 (<b>E</b>), TAp73 (<b>F</b>), ΔNp73 (<b>G</b>), and NME1 (<b>H</b>) were assessed 24 h after exposure to 0.5 mM CuSO<sub>4</sub> and/or 1, 5, and 10 µg/mL myricitrin. Total cellular proteins were extracted, separated by polyacrylamide gel electrophoresis, and transferred to nitrocellulose membranes. The membranes were incubated with specific primary antibodies, and subsequently with horseradish peroxidase-conjugated secondary antibodies. Protein bands were visualized using enhanced chemiluminescence. β-actin was used as a loading control to normalize protein expression levels. Data are presented as mean ± SD from 3 to 4 independent experiments. Densitometric analysis was performed using ImageJ (Fiji) software. Data were analyzed using an unpaired <span class="html-italic">t</span>-test (* <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01 vs. copper group; <sup>a</sup> <span class="html-italic">p</span> &lt; 0.05, <sup>b</sup> <span class="html-italic">p</span> &lt; 0.01, and <sup>c</sup> <span class="html-italic">p</span> &lt; 0.001 vs. control group). Representative Western blot images are shown in (<b>I</b>).</p>
Full article ">Figure 5
<p>Effects of specific inhibitors and modulators of intracellular signalling pathways on the neurotoxic effect of myricitrin. Cells were treated simultaneously with 0.5 mM CuSO<sub>4</sub> and 20 µg/mL myricitrin in combination with the following inhibitors: pifithrin-α (p53 inhibitor) (<b>A</b>), PJ34 (PARP inhibitor) (<b>B</b>), wortmannin (PI3K/Akt inhibitor) (<b>C</b>), U0126 (ERK1/2 inhibitor) (<b>D</b>), SB203580 (p38 inhibitor) (<b>E</b>), SP600125 (JNK inhibitor) (<b>F</b>), BAPTA-AM (intracellular calcium chelator) (<b>G</b>), MK-801 (NMDA receptor channel blocker) (<b>H</b>), nifedipine L-type calcium channel inhibitor) (<b>I</b>), and leupeptin (calpain inhibitor) (<b>J</b>). Inhibitors were applied 1 h prior to and during the 24 h treatment with copper and myricitrin. Data are presented as means ± SD from 3 to 5 independent experiments performed in triplicate. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and **** <span class="html-italic">p</span> &lt; 0.0001 vs. copper + myricitrin treatment (one-way ANOVA followed by post hoc Tukey’s test).</p>
Full article ">Figure 6
<p>2D height images of SH-SY5Y neuroblastoma cells using AFM: (<b>A</b>) control cell; (<b>B</b>) cell treated with 10 μg/mL myricitrin for 24 h; (<b>C</b>) cell treated with copper (0.5 mM CuSO<sub>4</sub>) for 24 h; (<b>D</b>) cell treated with myricitrin and copper (same concentrations as in the individual treatment) for 24 h. The scan area is 2 μm × 2 μm, with vertical scales noted on the images.</p>
Full article ">Figure 7
<p>2D height images (<b>left</b>) and height section profiles along the yellow line on the height image (<b>right</b>) of neuroblastoma SH-SY5Y cells using AFM: (<b>A</b>) control cell without treatment; (<b>B</b>) cell treated with 10 μg/mL myricitrin for 24 h; (<b>C</b>) cell treated with copper (0.5 mM CuSO<sub>4</sub>) for 24 h; (<b>D</b>) cell treated with myricitrin and copper (same concentrations as in the individual treatment) for 24 h.</p>
Full article ">Figure 8
<p>3D height images (<b>left</b>) and height section profiles along the yellow line on the height image (<b>right</b>) of neuroblastoma SH-SY5Y cells using AFM: (<b>A</b>) control cell without treatment; (<b>B</b>) cell treated with 10 μg/mL myricitrin for 24 h; (<b>C</b>) cell treated with copper (0.5 mM CuSO<sub>4</sub>) for 24 h; (<b>D</b>) cell treated with myricitrin and copper (same concentrations as in the individual treatment) for 24 h.</p>
Full article ">Figure 9
<p>Elasticity histograms of neuroblastoma SH-SY5Y cells from two independent experiments: (<b>A</b>) control cells, (<span class="html-italic">N</span><sub>cell</sub> = 10), <span class="html-italic">N</span><sub>fc</sub> = 580; (<b>B</b>) cell treated with 10 μg/mL myricitrin for 24 h, (<span class="html-italic">N</span><sub>cell</sub> = 9), <span class="html-italic">N</span><sub>fc</sub> = 580; (<b>C</b>) cells treated with copper (0.5 mM CuSO<sub>4</sub>) for 24 h, (<span class="html-italic">N</span><sub>cell</sub> = 12), <span class="html-italic">N</span><sub>fc</sub> = 580; (<b>D</b>) cells treated with myricitrin and copper for 24 h, (<span class="html-italic">N</span><sub>cell</sub> = 7), <span class="html-italic">N</span><sub>fc</sub> = 580. Histograms were fitted with the Gauss function, labelled by red and black lines for each measurement.</p>
Full article ">
13 pages, 2099 KiB  
Article
An Experimental and Quantum Chemical Calculation Study on the Performance of Different Types of Ester Collectors
by Di Wu, Jianhua Chen and Yuqiong Li
Molecules 2025, 30(1), 147; https://doi.org/10.3390/molecules30010147 - 2 Jan 2025
Viewed by 315
Abstract
Ester collectors have rapidly developed into the main flotation collectors for copper sulfide minerals since they were developed. In this study, the collecting performance of four collectors, O-isopropyl-N-ethyl thionocarbamate ester (IPETC), 3-pentyl xanthate acrylate ester (PXA), O-isobutyl-N-allyl-thionocarbamate (IBALTC), and O-isobutyl-N-isobutoxycarbonyl-thionocarbamate (IBIBCTC), was investigated [...] Read more.
Ester collectors have rapidly developed into the main flotation collectors for copper sulfide minerals since they were developed. In this study, the collecting performance of four collectors, O-isopropyl-N-ethyl thionocarbamate ester (IPETC), 3-pentyl xanthate acrylate ester (PXA), O-isobutyl-N-allyl-thionocarbamate (IBALTC), and O-isobutyl-N-isobutoxycarbonyl-thionocarbamate (IBIBCTC), was investigated through microflotation tests, microcalorimetric measurements, and quantum chemical calculations. The results of the microflotation tests show that IBALTC and IPETC have stronger collecting abilities than IBIBCTC and PXA; the order of collecting ability is IBALTC > IPETC > IBIBCTC > PXA. The microcalorimetry test also shows that the adsorption heat of the former two is higher. Quantum chemical calculations show the energy difference between the HOMOs of the collector and the LUMOs of minerals. The electrostatic potential extremum around S atom and the first ionization potential of IPETC and IBALTC are similar and were smaller than IBIBCTC and PXA, which shows that the collecting ability of the former two is similar and stronger than the latter two. Among the collectors, the S atom polarizability, electrophilic, and nucleophilic attack index of IBALTC are the largest, indicating that its electronic deformation capability and nucleophilic properties are the strongest, which results in the strongest coordination interaction with the copper ions in copper sulfide minerals and thus the highest collecting ability. The S atom polarizability, electrophilic, and nucleophilic attack index of PXA are the smallest, indicating that its electronic deformation capability and nucleophilicity are the weakest, and its collecting ability is the weakest. The coordination between collector and mineral surface was analyzed theoretically. The research results are of great help to the design and development of ester collectors. Full article
Show Figures

Figure 1

Figure 1
<p>Microflotation recoveries of chalcopyrite using IPETC, PXA, IBALTC, and IBIBCTC collectors.</p>
Full article ">Figure 2
<p>Heat flow of different collectors on chalcopyrite surface.</p>
Full article ">Figure 3
<p>Geometry of collectors’ molecular structures after optimization.</p>
Full article ">Figure 4
<p>Frontier orbital configurations of collector molecules.</p>
Full article ">Figure 5
<p>Electrostatic potential surface of collectors.</p>
Full article ">Figure 6
<p>Coordination model between surface Cu<sup>+</sup> and collectors.</p>
Full article ">Figure 7
<p>X-ray diffraction spectra of chalcopyrite.</p>
Full article ">
15 pages, 2216 KiB  
Article
Improved Voltammetric Procedure for Chloride Determination in Highly Acidic Media
by Rafał Maciąg, Wojciech Hyk, Tomasz Ratajczyk and Mikołaj Donten
Materials 2025, 18(1), 136; https://doi.org/10.3390/ma18010136 - 31 Dec 2024
Viewed by 337
Abstract
Cyclic voltammetry (CV) can be applied as a reliable method for the determination of chloride ions in a range from several to a couple hundred (about 200) ppm. Since the standard potential of chloride ion/gaseous chlorine is 1.36 V vs. normal hydrogen electrode [...] Read more.
Cyclic voltammetry (CV) can be applied as a reliable method for the determination of chloride ions in a range from several to a couple hundred (about 200) ppm. Since the standard potential of chloride ion/gaseous chlorine is 1.36 V vs. normal hydrogen electrode (NHE), the efficient oxidation of Cl ion occurs at very positive electrode potentials, usually higher than +1.7 V vs. NHE. It is possible to observe this phenomenon only at noble-metal or inert carbon electrodes. Many other solutes, usually organic compounds, are often oxidized at this potential. This is the reason why the determination of Cl content directly from an increase in the oxidation current is not reliable and could lead to overestimated values. However, gaseous chlorine, generated at a positive potential dissolve in the analyzed solution, could be reduced in the reverse scan of a cyclic voltammetric curve. Optimization of the experimental procedure using statistical tools enables the development of an improved method for the direct quantification of chloride ions in acid copper electroplating baths. Under the proposed experimental conditions, the calibration curve (Cl2 voltammetric reduction current vs. chloride ions concentration) is represented by the linear model for the concentration of chlorides ranging from 10 to 200 mg/dm3. The developed method for analyzing chloride ions in an acid sulfate electroplating copper bath has many unique properties. It is fast; the time of a single analysis is less than 20 min. In automatic mode, it can be repeated up to 50 times a day. The method does not require processing of the sample of the analyzed bath before measurement. As a result, no additional chemical reagents are used, and the test sample can be returned to the plating bath. Full article
(This article belongs to the Special Issue Electrochemical Material Science and Electrode Processes)
Show Figures

Figure 1

Figure 1
<p>A prototype of the voltammetric cell.</p>
Full article ">Figure 2
<p>OCP variation with time for a QRE (Cu<sup>2+</sup>/Cu) made from a sample of a newly prepared acid copper plating bath The QRE was tested in a newly prepared acid copper plating bath.</p>
Full article ">Figure 3
<p>OCP variation with time for a QRE (Cu<sup>2+</sup>/Cu) made from a sample of a production acid copper plating bath. The QRE was tested in the production acid copper plating bath too.</p>
Full article ">Figure 4
<p>Potential changes depend on variations in copper concentration and temperature, which were permissible within the operating range for DuPont’s ELECTROPOSIT™1300 Acid Copper technology.</p>
Full article ">Figure 5
<p>Cyclic voltammograms recorded at acid copper plating bath. Signals at about 1.28 V correspond to Cl<sup>−</sup> oxidation signals at about 1.08 V corresponds to Cl<sub>2</sub> reduction. All three CV curves were registered at sweep rates of 0.1 V/s for each direction of potential sweep. These voltammograms were recorded at a polycrystalline Pt electrode (temperature 298 K).</p>
Full article ">Figure 6
<p>The calibration curve for chloride concentration ranging from 10 to 200 mg/dm<sup>3</sup>. The colors of the curves correspond to the following chloride ion concentrations: dark blue—10 mg/dm<sup>3</sup>; orange—12 mg/dm<sup>3</sup>; gray—20 mg/dm<sup>3</sup>; yellow—40 mg/dm<sup>3</sup>; blue—60 mg/dm<sup>3</sup>; green—100 mg/dm<sup>3</sup>; navy blue—150 mg/dm<sup>3</sup>, brown—200 mg/dm<sup>3</sup>.</p>
Full article ">
22 pages, 13981 KiB  
Article
Simulation of Arc Discharge in an Argon/Methane Mixture, Taking into Account the Evaporation of Anode Material in Problems Related to the Synthesis of Functional Nanostructures
by Almaz Saifutdinov and Boris Timerkaev
Nanomaterials 2025, 15(1), 54; https://doi.org/10.3390/nano15010054 - 31 Dec 2024
Viewed by 416
Abstract
In this work, within the framework of a self-consistent model of arc discharge, a simulation of plasma parameters in a mixture of argon and methane was carried out, taking into account the evaporation of the electrode material in the case of a refractory [...] Read more.
In this work, within the framework of a self-consistent model of arc discharge, a simulation of plasma parameters in a mixture of argon and methane was carried out, taking into account the evaporation of the electrode material in the case of a refractory and non-refractory cathode. It is shown that in the case of a refractory tungsten cathode, almost the same methane conversion rate is observed, leading to similar values in the density of the main methane conversion products (C, C2, H) at different values of the discharge current density. However, with an increase in the current density, the evaporation rate of copper atoms from the anode increases, and a jump in the IV characteristic is observed, caused by a change in the plasma-forming ion. This is due to the lower ionization energy of copper atoms compared to argon atoms. In this mode, an increase in metal–carbon nanoparticles is expected. It is shown that, in the case of a cathode made of non-refractory copper, the discharge characteristics and the component composition of the plasma depend on the field enhancement factor near the cathode surface. It is demonstrated that increasing the field enhancement factor leads to more efficient thermal field emission, lowering the cathode’s surface temperature and the gas temperature in the discharge gap. This leads to the fact that, in the arc discharge mode with a cathode made of non-refractory copper, the dominant types of particles from which the synthesis of a nanostructure can begin are, in descending order, copper atoms (Cu), carbon clusters (C2), and carbon atoms (C). Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Electron spectrum in metal and different types of emissions.</p>
Full article ">Figure 2
<p>Schematic diagram of an arc discharge with the arrangement of electrodes (<b>top</b>) and the calculation area (<b>bottom</b>). The length of the electrodes is <span class="html-italic">L</span> and the length of the discharge gap is <span class="html-italic">L<sub>gap</sub></span>.</p>
Full article ">Figure 3
<p>Dependence of (<b>a</b>) voltage, surface temperature of the cathode and anode, (<b>b</b>) averaged densities of dominant types of particles, and their ions over the discharge gap on current density.</p>
Full article ">Figure 4
<p>Dynamics of changes in the average densities of different types of neutral particles over the discharge gap at different values of discharge current, (<b>a</b>) <span class="html-italic">I</span> = 10 A and (<b>b</b>) <span class="html-italic">I</span> = 60 A.</p>
Full article ">Figure 5
<p>Dynamics of changes in the average densities of different types of charged particles over the discharge gap at different values of discharge current, (<b>a</b>) <span class="html-italic">I</span> = 10 A and (<b>b</b>) <span class="html-italic">I</span> = 60 A.</p>
Full article ">Figure 6
<p>Distributions of charged particle densities along the discharge gap (the cathode is on the left and the anode is on the right) for two values of current density, (<b>a</b>) <span class="html-italic">j</span> = 7.8 × 10<sup>5</sup> A/m<sup>2</sup> (<span class="html-italic">I</span> = 10 A) and (<b>b</b>) <span class="html-italic">j</span> = 4.8 × 10<sup>6</sup> A/m<sup>2</sup> (<span class="html-italic">I</span> = 60A).</p>
Full article ">Figure 7
<p>Dependences of (<b>a</b>) the voltage across the discharge gap and (<b>b</b>) the surface temperature of the cathode and anode on the current density for different values of the field enhancement factor near the cathode.</p>
Full article ">Figure 8
<p>Dependences of the averaged densities of the dominant types of neutral and charged particles over the discharge gap as a function of the current density for different values of the field enhancement factor on the cathode surface: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math>, (<b>b)</b> <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>250</mn> </mrow> </semantics></math>, (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>300</mn> </mrow> </semantics></math>, and (<b>d</b>) <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Distributions of densities of neutral and charged particles along the discharge gap (the cathode is on the left and the anode is on the right) at a current density of <span class="html-italic">j</span> = 7.8 × 10<sup>5</sup> A/m<sup>2</sup> for two values of the field enhancement factor, (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Distributions of (<b>a</b>) the temperature of heavy components of plasma, (<b>b</b>) the temperature of electrons and (<b>c</b>) the electric potential at a current density of <span class="html-italic">j</span> = 7.8 × 10<sup>5</sup> A/m<sup>2</sup> for two values of the field enhancement factor, <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mi>β</mi> <mo>=</mo> <mn>400</mn> </mrow> </semantics></math>.</p>
Full article ">
13 pages, 3997 KiB  
Article
Reliable Atom Probe Tomography of Cu Nanoparticles Through Tailored Encapsulation by an Electrodeposited Film
by Aydan Çiçek, Florian Knabl, Maximilian Schiester, Helene Waldl, Lidija D. Rafailović, Michael Tkadletz and Christian Mitterer
Nanomaterials 2025, 15(1), 43; https://doi.org/10.3390/nano15010043 - 30 Dec 2024
Viewed by 439
Abstract
Nanoparticles are essential for energy storage, catalysis, and medical applications, emphasizing their accurate chemical characterization. However, atom probe tomography (APT) of nanoparticles sandwiched at the interface between an encapsulating film and a substrate poses difficulties. Poor adhesion at the film-substrate interface can cause [...] Read more.
Nanoparticles are essential for energy storage, catalysis, and medical applications, emphasizing their accurate chemical characterization. However, atom probe tomography (APT) of nanoparticles sandwiched at the interface between an encapsulating film and a substrate poses difficulties. Poor adhesion at the film-substrate interface can cause specimen fracture during APT, while impurities may introduce additional peaks in the mass spectra. We demonstrate preparing APT specimens with strong adhesion between nanoparticles and film/substrate matrices for successful analysis. Copper nanoparticles were encapsulated at the interface between nickel film and cobalt substrate using electrodeposition. Cobalt and nickel were chosen to match their evaporation fields with copper, minimizing peak overlaps and aiding nanoparticle localization. Copper nanoparticles were deposited via magnetron sputter inert gas condensation with varying deposition times to yield suitable surface coverages, followed by encapsulation with the nickel film. In-plane and cross-plane APT specimens were prepared by femtosecond laser ablation and focused ion beam milling. Longer deposition times resulted in agglomerated nanoparticles as well as pores and voids, causing poor adhesion and specimen failure. In contrast, shorter deposition times provided sufficient surface coverage, ensuring strong adhesion and reducing void formation. This study emphasizes controlled surface coverage for reliable APT analysis, offering insights into nanoparticle chemistry. Full article
(This article belongs to the Section Synthesis, Interfaces and Nanostructures)
Show Figures

Figure 1

Figure 1
<p>Schematic of NP synthesis by DC magnetron sputter inert gas condensation.</p>
Full article ">Figure 2
<p>Temporal evolution of the grid current over NP size recorded in-situ during two separate deposition runs using the QMF.</p>
Full article ">Figure 3
<p>Secondary electron SEM micrographs of Cu NPs on Co substrates after deposition times of (<b>a</b>) 30 min and (<b>b</b>) 2 min. (<b>c</b>) Corresponding size distributions of Cu NPs, where the obtained histograms with a bin width of 5 nm have been fitted with Gaussian functions.</p>
Full article ">Figure 4
<p>(<b>a</b>) X-ray diffractogram for the specimen containing Cu NPs (deposition time 2 min) encapsulated between Ni film and Co substrate. Reference peak positions of fcc-Ni and hcp-Co were taken from Ref. [<a href="#B37-nanomaterials-15-00043" class="html-bibr">37</a>]. (<b>b</b>) Corresponding SEM image of broad ion beam polished cross-section of the specimen. The regions marked with (<b>d</b>,<b>e</b>) highlight in-plane and cross-plane specimen preparation for the subsequent APT measurement. (<b>c</b>) SEM image of a FIB cross-section with Cu NPs (deposition time 30 min) encapsulated between the Ni film and the Co substrate, where region (<b>f</b>) indicates cross-plane APT specimen preparation. Yellow arrows in the pre-prepared APT specimens in (<b>d</b>–<b>f</b>) mark the interfaces between Co substrate and Ni encapsulation film.</p>
Full article ">Figure 5
<p>Summary of the outcome of the APT investigation of a cross-plane specimen with Cu NPs (deposition time 30 min) encapsulated between Ni film and Co substrate, with 6 million ions detected during the measurement: (<b>a</b>) Clipped 3D APT reconstruction with a cylindrical volume of 20 × 20 × 15 nm<sup>3</sup> showing the interface in ROI 1, where Cu NPs are encapsulated (the isoconcentration surface of Cu NPs was set at 2 at%). (<b>b</b>) Corresponding 1D concentration profiles along the axis of the cylindrical ROI 1 indicated in (<b>a</b>). (<b>c</b>) Corresponding mass spectrum in the 58–65 Da region of ROI 1. (<b>d</b>) Side-view clipped APT reconstruction with a cylindrical volume of 23 × 45 × 45 nm<sup>3</sup>, representing ROI 2 positioned at the fractured interface. (<b>e</b>) 2D contour plot of the Cu concentration and (<b>f</b>) top view of ROI 2 showing the spatial distribution of Cu atoms towards the z axis of the reconstruction. Note that the spheres shown in (<b>f</b>) are used solely for visualization of a fraction of the detected Cu ions and should not be interpreted as individual NPs.</p>
Full article ">Figure 6
<p>Summary of the outcome of the APT investigation of an in-plane specimen with Cu NPs (deposition time 2 min) encapsulated between Ni film and Co substrate, with 40 million ions detected during the measurement. (<b>a</b>) Clipped 3D APT reconstruction with a cylindrical volume of 20 × 20 × 50 nm³ showing the interface in ROI 3, where Cu NPs are encapsulated (the isoconcentration surface of Cu NPs was set at 2 at%). (<b>b</b>) Corresponding 1D concentration profiles along the axis of the cylindrical ROI 3 indicated in (<b>a</b>). (<b>c</b>) Corresponding mass spectrum in the 58–65 Da region of the ROI 3. (<b>d</b>) Side-view clipped APT reconstruction with a cylindrical volume of 10 × 10 × 30 nm<sup>3</sup> representing ROI 4. (<b>e</b>) 2D contour plot of the Cu concentration and (<b>f</b>) top view of ROI 4 showing the spatial distribution of Cu NPs towards the direction indicated by the black arrow in (<b>d</b>). Note that the spheres shown in (<b>f</b>) are used solely for visualization of a fraction of the detected Cu ions and should not be interpreted as individual NPs.</p>
Full article ">Figure 7
<p>Summary of the outcome of the APT investigation of a cross-plane specimen with Cu NPs (deposition time 2 min) encapsulated between Ni film and Co substrate, with 4.5 million ions detected during the measurement. (<b>a</b>) Clipped 3D reconstruction with a cylindrical volume of 20 × 20 × 15 nm³ showing the interface of the frontside of the APT specimen in region of interest 5 (ROI 5), where Cu NPs are encapsulated (the isoconcentration surface of Cu NPs was set at 2 at%). (<b>b</b>) Corresponding 1D concentration profiles along the cylindrical ROI 5 indicated in (<b>a</b>). (<b>c</b>) Corresponding mass spectrum of Cu peaks in the 58–65 Da region of the ROI 5. (<b>d</b>) Side-view of the backside of the clipped APT specimen with a cylindrical volume of 10 × 30 × 30 nm<sup>3</sup> representing ROI 6 positioned at the interface. (<b>e</b>) 2D contour plot of the Cu concentration and (<b>f</b>) top view of ROI 6 showing the spatial distribution of Cu NPs towards the direction indicated by the black arrow in (<b>d</b>). Note that the spheres shown in (<b>f</b>) are used solely for visualization of a fraction of the detected Cu ions and should not be interpreted as representing individual NPs.</p>
Full article ">
12 pages, 1974 KiB  
Article
The Effects of Polystyrene Microplastics and Copper Ion Co-Contamination on the Growth of Rice Seedlings
by Huiyu Jin, Guohe Lin, Mingzi Ma, Lin Wang and Lixiang Liu
Nanomaterials 2025, 15(1), 17; https://doi.org/10.3390/nano15010017 - 26 Dec 2024
Viewed by 350
Abstract
Microplastics (MPs) are emerging pollutants of global concern, while heavy metals such as copper ions (Cu2+) are longstanding environmental contaminants with well-documented toxicity. This study investigates the independent and combined effects of polystyrene microplastics (PS-MPs) and Cu on the physiological and [...] Read more.
Microplastics (MPs) are emerging pollutants of global concern, while heavy metals such as copper ions (Cu2+) are longstanding environmental contaminants with well-documented toxicity. This study investigates the independent and combined effects of polystyrene microplastics (PS-MPs) and Cu on the physiological and biochemical responses of rice seedlings (Oryza sativa L.), a key staple crop. Hydroponic experiments were conducted under four treatment conditions: control (CK), PS-MPs (50 mg·L−1), Cu (20 mg·L−1 Cu2+), and a combined PS-MPs + Cu treatment. The results showed that PS-MPs had a slight stimulatory effect on root elongation, while Cu exposure mildly inhibited root growth. However, the combined treatment (PS-MPs + Cu) demonstrated no significant synergistic or additive toxicity on growth parameters such as root, stem, and leaf lengths or biomass (fresh and dry weights). Both PS-MPs and Cu significantly reduced peroxidase (POD) activity in root, stem, and leaf, indicating oxidative stress and disrupted antioxidant defenses. Notably, in the combined treatment, PS-MPs mitigated Cu toxicity by adsorbing Cu2+ ions, reducing their bioavailability, and limiting Cu accumulation in rice tissues. These findings reveal a complex interaction between MPs and heavy metals in agricultural systems. While PS-MPs can alleviate Cu toxicity by reducing its bioavailability, they also compromise antioxidant activity, potentially affecting plant resilience to stress. This study provides a foundation for understanding the combined effects of MPs and heavy metals, emphasizing the need for further research into their long-term environmental and agronomic impacts. Full article
(This article belongs to the Section Environmental Nanoscience and Nanotechnology)
Show Figures

Figure 1

Figure 1
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on root length of rice seedlings. The letter of a indicates insignificant differences between treatments.</p>
Full article ">Figure 2
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on shoot length of rice seedlings. The letter of a indicates insignificant differences between treatments.</p>
Full article ">Figure 3
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on fresh weight of rice seedlings. The letter of a indicates insignificant differences between treatments.</p>
Full article ">Figure 4
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on dry weight of rice seedlings. The letter of a indicates insignificant differences between treatments.</p>
Full article ">Figure 5
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on root POD activity of rice seedlings. Different letters a, b, c and d represent significant difference between different treatments (at <span class="html-italic">p</span> ≤ 0.05) whereas, same letters indicate insignificant differences between treatments.</p>
Full article ">Figure 6
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on shoot POD activity of rice seedlings. Different letters a, b and c represent significant difference between different treatments (at <span class="html-italic">p</span> ≤ 0.05) whereas, same letters indicate insignificant differences between treatments.</p>
Full article ">Figure 7
<p>Effect of single and combined treatment of PS-MP and Cu<sup>2+</sup> on Cu content of rice seedlings. Different letters a, b and c represent significant difference between different treatments (at <span class="html-italic">p</span> ≤ 0.05) whereas, same letters indicate insignificant differences between treatments.</p>
Full article ">
14 pages, 11164 KiB  
Article
Photoelectron Spectroscopy Study of the Optical and Electrical Properties of Cr/Cu/Mn Tri-Doped Bismuth Niobate Pyrochlore
by Nadezhda A. Zhuk, Nikolay A. Sekushin, Maria G. Krzhizhanovskaya, Artem A. Selutin, Aleksandra V. Koroleva, Ksenia A. Badanina, Sergey V. Nekipelov, Olga V. Petrova and Victor N. Sivkov
Sci 2025, 7(1), 1; https://doi.org/10.3390/sci7010001 - 26 Dec 2024
Viewed by 411
Abstract
The multielement pyrochlore of the composition Bi1.57Mn1/3Cr1/3Cu1/3Nb2O9−Δ (sp. gr. Fd-3m:2, 10.4724 Å) containing transition element atoms—chromium, manganese and copper in equimolar amounts—was synthesized for the first time using the solid-phase reaction method. [...] Read more.
The multielement pyrochlore of the composition Bi1.57Mn1/3Cr1/3Cu1/3Nb2O9−Δ (sp. gr. Fd-3m:2, 10.4724 Å) containing transition element atoms—chromium, manganese and copper in equimolar amounts—was synthesized for the first time using the solid-phase reaction method. The microstructure of the ceramics is grainless and has low porosity. The sample is characterized by reflection in the red (705 nm) color region. The band gap for the direct allowed transition in the sample is 1.68 eV. The parameters of the Bi5d, Nb3d, Сr2p, Mn2p, and Cu2p X-ray photoelectron spectroscopy (XPS) spectra for the mixed pyrochlore are compared with the parameters of transition element oxides. For the complex pyrochlore, a characteristic shift in the Bi4f and Nb3d spectra to the region of lower energies by 0.15 and 0.60 eV, respectively, is observed. According to the XPS Cu2p and Mn2p spectra of pyrochlore, copper, and manganese cations are in a mixed charge state; they mainly have an effective charge of +2/+3, and the Cr2p spectrum is a superposition of the spectra of chromium ions in the charge state of +3, +4, +6. At 24 °С, the permittivity of the sample in the frequency range (104–106 Hz) weakly depends on the frequency and is equal to ~100, the dielectric loss tangent is 0.017. The activation energy of conductivity is equal to 0.41 eV. The specific electrical conductivity of Bi1.57Cr1/3Cu1/3Mn1/3Nb2O9−Δ increases with the temperature increasing from 1.8 × 10−5 Ohm−1·m−1 (24 °С) to 0.1 Ohm−1·m−1 (330 °С). Nyquist curves for the sample are modeled by equivalent electrical circuits. Full article
(This article belongs to the Section Chemistry Science)
Show Figures

Figure 1

Figure 1
<p>Experimental (blue circles), calculated (solid red line) and difference (grey line) XRD patterns of Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ.</sub> The agreement factors (R-factors) in Rietveld analysis are given.</p>
Full article ">Figure 2
<p>Micrographs of the surface of the Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> sample in the secondary (<b>a</b>) and elastically reflected (<b>b</b>) electron modes.</p>
Full article ">Figure 3
<p>Element maps and EDS spectrum for the Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> sample.</p>
Full article ">Figure 4
<p>Diffuse reflectance spectrum (<b>a</b>) and Tauc curve for Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> (<b>b</b>).</p>
Full article ">Figure 5
<p>Total XPS spectrum of Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> (BiCrCuMnNbO) (<b>a</b>); XPS Bi4f spectra in BiCrCuMnNbO. For comparison, the spectrum of Bi<sub>2</sub>O<sub>3</sub> is shown (<b>b</b>); XPS spectra of niobium and bismuth oxides and cations in BiCrCuMnNbO (<b>c</b>); XPS spectra of manganese oxides and Mn2p spectra in BiCrCuMnNbO (<b>d</b>); XPS Сr2p spectra in BiCrCuMnNbO and chromium oxides (<b>e</b>); XPS Cu2p spectra in BiCrCuMnNbO; Cu<sub>2</sub>O and CuO (<b>f</b>).</p>
Full article ">Figure 6
<p>The modulus (<b>a</b>) and phase of the impedance (<b>b</b>) of the Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> sample.</p>
Full article ">Figure 7
<p>The dielectric permittivity (<b>a</b>) and tangent of dielectric losses (<b>b</b>) in the frequency range 25–(5 × 10<sup>6</sup>) Hz at 24–250 °С of the Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub>; the temperature dependence of the end-to-end conductivity of the sample, plotted on the Arrhenius scale for the temperature range of 24–350 °С (<b>c</b>).</p>
Full article ">Figure 8
<p>Hodographs of impedance of the Bi<sub>1.57</sub>Cr<sub>1/3</sub>Cu<sub>1/3</sub>Mn<sub>1/3</sub>Nb<sub>2</sub>O<sub>9−Δ</sub> sample, measured at temperatures 24–250 °С.</p>
Full article ">
18 pages, 6483 KiB  
Article
Surface Chemistry Aspects of Ion Exchange in Basic Copper Salts
by Sebastian Skupiński, Marta Kalbarczyk, Daniel Kamiński and Marek Kosmulski
Molecules 2025, 30(1), 21; https://doi.org/10.3390/molecules30010021 - 25 Dec 2024
Viewed by 289
Abstract
Brochantite was precipitated using stoichiometric amounts of CuSO4 and NaOH and characterized by scanning electron microscopy, specific surface area, thermogravimetric analysis, and zeta potential. Brochantite can be converted into paratacamite, basic copper bromide, and copper phthalate by shaking the powder with solutions [...] Read more.
Brochantite was precipitated using stoichiometric amounts of CuSO4 and NaOH and characterized by scanning electron microscopy, specific surface area, thermogravimetric analysis, and zeta potential. Brochantite can be converted into paratacamite, basic copper bromide, and copper phthalate by shaking the powder with solutions containing excess corresponding anions. By contrast, attempts to convert brochantite into basic iodide, acetate, nitrate, or rhodanide in a similar way failed, that is, the powder after shaking with solutions containing excess corresponding anions still showed the powder X-ray diffraction pattern of brochantite. Successful ion exchange resulted in a decrease in the specific surface area by a factor of 10, but the specific surface area was unchanged when attempts to exchange the anion failed. Interestingly enough, paratacamite can also be converted into brochantite by shaking with solution containing excess sulfate. Brochantite and paratacamite obtained by precipitation and the salts obtained by ion exchange showed a negative zeta potential at pH > 9. Full article
Show Figures

Figure 1

Figure 1
<p>TGA curves of original brochantite (<b>A</b>) and paratacamite (<b>B</b>).</p>
Full article ">Figure 2
<p>SEM image of brochantite.</p>
Full article ">Figure 3
<p>SEM image of paratacamite.</p>
Full article ">Figure 4
<p>XRD pattern of brochantite. Line: experimental. Bars: from RRUFF repository.</p>
Full article ">Figure 5
<p>XRD pattern of paratacamite. Line: experimental. Bars: from RRUFF repository.</p>
Full article ">Figure 6
<p>Electrokinetic curves of the original brochantite.</p>
Full article ">Figure 7
<p>XRD patterns of B003, B005, B006 and B010. Lines: experimental. Bars: brochantite from RRUFF repository.</p>
Full article ">Figure 8
<p>XRD patterns of B001, B002, B007, and B008. Lines: experimental. Bars: brochantite from RRUFF repository.</p>
Full article ">Figure 9
<p>XRD patterns of B001, B008, and B012. Lines: experimental. Bars: atacamite (white) and paratacamite (black) from RRUFF repository.</p>
Full article ">Figure 10
<p>XRD patterns of B004, B014, and B015. Lines: experimental. Bars: brochantite from RRUFF repository.</p>
Full article ">Figure 11
<p>XRD patterns of B011, P017, and P018. Lines: experimental. Bars: brochantite (green) and paratacamite (black) from RRUFF repository.</p>
Full article ">Figure 12
<p>XRD patterns of P013 and P016. Lines: experimental. Bars: brochantite from RRUFF repository.</p>
Full article ">Figure 13
<p>Electrokinetic curves of B003, B005, B006, and B010.</p>
Full article ">Figure 14
<p>Electrokinetic curves of B001, B002, B007, B008, and B012.</p>
Full article ">Figure 15
<p>Electrokinetic curves of B004, B008, B014, and B015.</p>
Full article ">Figure 16
<p>Electrokinetic curves of B011, P017, and P018.</p>
Full article ">Figure 17
<p>Electrokinetic curves of P013 and P016.</p>
Full article ">
16 pages, 3520 KiB  
Article
Low Temperature Synthesis of 3d Metal (Fe, Co, Ni, Cu)-Doped TiO2 Photocatalyst via Liquid Phase Deposition Technique
by Mitsuhiro Honda, Yusaku Yoshii, Nobuchika Okayama and Yo Ichikawa
Sustain. Chem. 2025, 6(1), 1; https://doi.org/10.3390/suschem6010001 - 24 Dec 2024
Viewed by 476
Abstract
The titanium dioxide (TiO2) photocatalyst is an important semiconducting material that exhibits environmental purification functions when exposed to light. Elemental doping of TiO2 is considered an important strategy to improve its photocatalytic activity. Herein, we have achieved the low-temperature, atmospheric-pressure [...] Read more.
The titanium dioxide (TiO2) photocatalyst is an important semiconducting material that exhibits environmental purification functions when exposed to light. Elemental doping of TiO2 is considered an important strategy to improve its photocatalytic activity. Herein, we have achieved the low-temperature, atmospheric-pressure synthesis of anatase TiO2 particles with doping of 3d metals (Fe, Co, Ni and Cu) based on the liquid phase deposition technique. All products prepared by adding 3d metals were found to consist of TiO2 crystals in the anatase phase with a fine protruding structure of about 40 nm on the surface, as was the case without the addition of metal ions. Iron and copper were observed to be incorporated at higher concentrations than cobalt and nickel, with an elemental addition of up to 4 at% and 1 at%, respectively, when 10 mM iron and copper nitrate were applied. Such doping efficiency could be explained by the difference in ionic radius and chemical stability. A narrowing of the optical band gap with doping elements was also observed, and it was found that optical sensitivity could be imparted down to the visible-light region of 2.4 eV (Fe: 4 at% addition). Furthermore, the 3d metal-doped TiO2 demonstrated in this study was shown to exhibit photocatalytic methane degradation activity. The amount of methane degradation per unit area of the microparticles was twice as great when iron and copper were added, compared to the undoped counterpart. It has been demonstrated that the strategy of doping TiO2 with 3d metal ions by low-temperature synthesis methods is effective in enhancing carrier dynamics and introducing surface active sites, thus increasing methane degradation activity. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of the setup for testing photocatalysis to degrade methane. (Inset: picture of the setup).</p>
Full article ">Figure 2
<p>SEM images of the products prepared with doping Fe (<b>a</b>–<b>c</b>), Co (<b>d</b>–<b>f</b>), Ni (<b>g</b>–<b>i</b>), and Cu (<b>j</b>–<b>l</b>). The scale bar corresponds to 200 nm.</p>
Full article ">Figure 3
<p>Raman spectra of the products prepared with doping of Fe (orange), Co (red), Ni (blue), and Cu (green). Undoped TiO<sub>2</sub> (LPD, commercial) is displayed in black color. E<sub>g</sub>, B<sub>1g</sub> and A<sub>1g</sub> indicate vibrational modes, which are derived from anatase TiO<sub>2</sub>.</p>
Full article ">Figure 4
<p>Molar concentration dependence on FWHM values of E<sub>g</sub> mode. Orange, blue, red and green colors indicate Fe, Ni, Co and Cu doping, respectively.</p>
Full article ">Figure 5
<p>Molar concentration dependence on peak wavenumber of E<sub>g</sub> mode. Orange, blue, red and green colors indicate Fe, Ni, Co and Cu doping, respectively.</p>
Full article ">Figure 6
<p>Atomic concentration of dopant elements in the product depends on the molar concentration. Orange, blue, red and green colors indicate Fe, Ni, Co and Cu doping, respectively.</p>
Full article ">Figure 7
<p>Correlation between bandgaps and molar concentrations. Orange, blue, red and green colors indicate Fe, Ni, Co and Cu doping, respectively.</p>
Full article ">Figure 8
<p>Reduction rate of methane with TiO<sub>2</sub> doping Fe, Co, Ni and Cu. The initial concentration of methane gas was (3600 ± 100) ppm. Photocatalytic reaction was initiated by illuminating 24 mg of photocatalyst powder entirely with black light (wavelength: 360 nm, power: (1.60 ± 0.05) mW/cm<sup>2</sup>).</p>
Full article ">Figure 9
<p>Rate of methane reduction per unit area. Photocatalytic reaction was initiated by illuminating the photocatalyst powder (24 mg) entirely with black light (wavelength: 360 nm, power: (1.60 ± 0.05) mW/cm<sup>2</sup>), while the methane reduction rate per unit area was calculated by dividing the reduction rate by the surface area.</p>
Full article ">Figure 10
<p>Narrow scans for O1s (<b>a</b>) and Ti2p (<b>b</b>) peaks for undoped and doped TiO<sub>2</sub>. The concentration of solutes applied corresponds to 0.1 mM. Dotted spectra are Voigt functions obtained through the deconvolution of spectra.</p>
Full article ">Figure 11
<p>Narrow scans for doped elements: (<b>a</b>) Fe2p, (<b>b</b>) Co2p, (<b>c</b>) Ni2p, and (<b>d</b>) Cu2p. Colors are prescribed to spectra of the samples in the same manner as those in <a href="#suschem-06-00001-f003" class="html-fig">Figure 3</a>, <a href="#suschem-06-00001-f004" class="html-fig">Figure 4</a>, <a href="#suschem-06-00001-f005" class="html-fig">Figure 5</a>, <a href="#suschem-06-00001-f006" class="html-fig">Figure 6</a> and <a href="#suschem-06-00001-f007" class="html-fig">Figure 7</a>. The applied concentration corresponds to 0.1 mM.</p>
Full article ">
23 pages, 2443 KiB  
Article
Neuroprotective Potential of Indole-Based Compounds: A Biochemical Study on Antioxidant Properties and Amyloid Disaggregation in Neuroblastoma Cells
by Tania Ciaglia, Maria Rosaria Miranda, Simone Di Micco, Mariapia Vietri, Gerardina Smaldone, Simona Musella, Veronica Di Sarno, Giulia Auriemma, Carla Sardo, Ornella Moltedo, Giacomo Pepe, Giuseppe Bifulco, Carmine Ostacolo, Pietro Campiglia, Michele Manfra, Vincenzo Vestuto and Alessia Bertamino
Antioxidants 2024, 13(12), 1585; https://doi.org/10.3390/antiox13121585 - 23 Dec 2024
Viewed by 521
Abstract
Based on the established neuroprotective properties of indole-based compounds and their significant potential as multi-targeted therapeutic agents, a series of synthetic indole–phenolic compounds was evaluated as multifunctional neuroprotectors. Each compound demonstrated metal-chelating properties, particularly in sequestering copper ions, with quantitative analysis revealing approximately [...] Read more.
Based on the established neuroprotective properties of indole-based compounds and their significant potential as multi-targeted therapeutic agents, a series of synthetic indole–phenolic compounds was evaluated as multifunctional neuroprotectors. Each compound demonstrated metal-chelating properties, particularly in sequestering copper ions, with quantitative analysis revealing approximately 40% chelating activity across all the compounds. In cellular models, these hybrid compounds exhibited strong antioxidant and cytoprotective effects, countering reactive oxygen species (ROS) generated by the Aβ(25–35) peptide and its oxidative byproduct, hydrogen peroxide, as demonstrated by quantitative analysis showing on average a 25% increase in cell viability and a reduction in ROS levels to basal states. Further analysis using thioflavin T fluorescence assays, circular dichroism, and computational studies indicated that the synthesized derivatives effectively promoted the self-disaggregation of the Aβ(25–35) fragment. Taken together, these findings suggest a unique profile of neuroprotective actions for indole–phenolic derivatives, combining chelating, antioxidant, and anti-aggregation properties, which position them as promising compounds for the development of multifunctional agents in Alzheimer’s disease therapy. The methods used provide reliable in vitro data, although further in vivo validation and assessment of blood–brain barrier penetration are needed to confirm therapeutic efficacy and safety. Full article
(This article belongs to the Section Health Outcomes of Antioxidants and Oxidative Stress)
Show Figures

Figure 1

Figure 1
<p>Scheme of the synthesized molecules.</p>
Full article ">Figure 2
<p>(<b>A</b>) UV spectra (in the range of 280 to 400 nm) of compounds (30 µM) alone and in the presence of 40 µM FeSO<sub>4</sub>, FeCl<sub>3</sub>, and CuSO<sub>4</sub>. (<b>B</b>) Copper-chelating quantitative analyses of compounds (30 µM). EDTA (1 mM) was used as the positive control. Results are shown as mean ± standard deviation (SD) from three independent experiments. **, *** denote <span class="html-italic">p</span> &lt; 0.01 and <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl.</p>
Full article ">Figure 3
<p>Neuroprotective activity of compounds. (<b>A</b>) SH-SY5Y cells were exposed to compounds at a concentration of 30 µM. Neuroprotective effects of compounds against (<b>B</b>) H<sub>2</sub>O<sub>2</sub>-induced (500 μM) cytotoxicity, (<b>C</b>) H<sub>2</sub>O<sub>2</sub>-induced (500 μM) ROS production, and (<b>D</b>) Aβ(25–35)-induced (40 μM) cytotoxicity. The 2′,7′-dichlorofluorescin diacetate (DCFH-DA) assay was conducted to reveal ROS production. The changes in viability were determined by calculating the percentage of viable cells in treated cultures relative to untreated controls. The results are presented as the mean ± standard deviation (SD) from three independent experiments. *, **, and *** denote, respectively, <span class="html-italic">p</span> &lt; 0.05, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl; <sup>##</sup>, and <sup>###</sup> denote, respectively, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001 vs. H<sub>2</sub>O<sub>2</sub> or Aβ(25–35).</p>
Full article ">Figure 4
<p>Disaggregating properties of compounds. In-cell ThT assay was performed for both fluorescence microscopy (<b>A</b>) and spectrophotometry (<b>B</b>). Scale bar: 100 μm. (<span class="html-italic">N</span> ≥ 10). Cells were observed at 20× magnification. (<b>C</b>) ThT shows a direct disaggregating effect of compounds against Aβ. Data are shown as mean ± SD of three different experiments performed in triplicate. *** denotes <span class="html-italic">p</span> &lt; 0.001 vs. Ctrl; <sup>##</sup>, and <sup>###</sup> denote, respectively, <span class="html-italic">p</span> &lt; 0.01, and <span class="html-italic">p</span> &lt; 0.001 vs. H<sub>2</sub>O<sub>2</sub> or Aβ(25–35).</p>
Full article ">Figure 5
<p>(<b>A</b>,<b>B</b>) CD curves and secondary structure analysis of the Aβ(25−35) peptide were performed using the CONTIN algorithm after 24 h of aggregation. Aβ(25–35) 40 μM was used as a positive control. A different colour has been selected for Aβ(25–35) and each compound in presence Aβ(25–35), as indicated in the legend.</p>
Full article ">Figure 6
<p>Three-dimensional model of the interactions of <b>12</b> (<b>A</b>), <b>13</b> (<b>B</b>), <b>14</b> (<b>C</b>), <b>20</b> (<b>D</b>), <b>21</b> (<b>E</b>), and <b>22</b> (<b>F</b>) with Aβ(25–35). The biological target is depicted by a ribbon colored by a chain (D, magenta; E, azure; F, green; G, faded red) and tube (colored: C, by chain; polar H, white; N, dark blue; O, red; S, yellow). The small molecules are represented by sticks (gray for <b>12</b>, cyan for <b>13</b>, violet for <b>14</b>, orange for <b>20</b>, black for <b>21</b>, khaki for <b>22</b>) and balls (colored: C, as for the sticks; polar H, white; N, dark blue; O, red). The hydrogen bonds between the ligand and Aβ(25–35) are represented by the dashed black lines.</p>
Full article ">
14 pages, 1356 KiB  
Article
Innovative Nafion- and Lignin-Based Cation Exchange Materials Against Standard Resins for the Removal of Heavy Metals During Water Treatment
by Sara Bergamasco, Luis Alexander Hein, Laura Silvestri, Robert Hartmann, Giampiero Menegatti, Alfonso Pozio and Antonio Rinaldi
Separations 2024, 11(12), 357; https://doi.org/10.3390/separations11120357 - 21 Dec 2024
Viewed by 485
Abstract
The contamination of water by heavy metals poses an escalating risk to human health and the environment, underscoring the critical need for efficient removal methods to secure safe water resources. This study evaluated the performance of four cationic exchange materials (labeled “PS—DVB”, “PA—DVB”, [...] Read more.
The contamination of water by heavy metals poses an escalating risk to human health and the environment, underscoring the critical need for efficient removal methods to secure safe water resources. This study evaluated the performance of four cationic exchange materials (labeled “PS—DVB”, “PA—DVB”, “TFSA”, and “OGL”) in removing or harvesting metals such as copper, silver, lead, cobalt, and nickel from aqueous solutions, several of which are precious and/or classified as Critical Raw Materials (CRMs) due to their economic importance and supply risk. The objective was to screen and benchmark the four ion exchange materials for water treatment applications by investigating their metal sequestration capacities. Experiments were conducted using synthetic solutions with controlled metal concentrations, analyzed through ICP-OES, and supported by kinetic modeling. The adsorption capacities (qe) obtained experimentally were compared with those predicted by pseudo-first-order and pseudo-second-order models. This methodology enables high precision and reproducibility, validating its applicability for assessing ion exchange performance. The results indicated that PS—DVB and PA—DVB resins proved to be of “wide range”, exhibiting high efficacy for most of the metals tested, including CRM-designated ones, and suggesting their suitability for water purification. Additionally, the second-life Nafion-based “TFSA” material demonstrated commendable performance, highlighting its potential as a viable and technologically advanced alternative in water treatment. Lastly, the lignin-based material, “OGL”, representing the most innovative and sustainability apt option, offered relevant performance only in selected cases. The significant differences in performance among the resins underscore the impact of structural and compositional factors on adsorption efficiency. This study offers valuable insights for investigating and selecting new sustainable materials for treating contaminated water, opening new pathways for targeted and optimized solutions in environmental remediation. Full article
(This article belongs to the Special Issue Separation Technology for Metal Extraction and Removal)
Show Figures

Figure 1

Figure 1
<p>Photographic images of the ion exchange materials utilized: (<b>A</b>) polystyrene–divinylbenzfene resin (PS–DVB), (<b>B</b>) polyacrylic–divinylbenzene resin (PA–DVB), (<b>D</b>) regenerated Nafion granules (TFSA), and (<b>D</b>) lignin-based powder material (OGL).</p>
Full article ">Figure 2
<p>Time-dependent concentration of the available metal ions (Cu<sup>2+</sup>, Pb<sup>2+</sup>, Co<sup>2+</sup> and Ni<sup>2+</sup>) in bulk solution (ppm) examined by ICP-EOS. Each panel shows the remaining concentration of a specific metal ion over time (0, 5, 10, 20, and 40 min): (<b>A</b>) Cu<sup>2+</sup> concentration in bulk, (<b>B</b>) Pb<sup>2+</sup> concentration in bulk, (<b>C</b>) Co<sup>2+</sup> concentration in bulk, and (<b>D</b>) Ni<sup>2+</sup> concentration in bulk.</p>
Full article ">
12 pages, 4468 KiB  
Article
Characterization of the Interaction of Human γS Crystallin with Metal Ions and Its Effect on Protein Aggregation
by Reinier Cardenas, Arline Fernandez-Silva, Vanesa Ramirez-Bello and Carlos Amero
Biomolecules 2024, 14(12), 1644; https://doi.org/10.3390/biom14121644 - 21 Dec 2024
Viewed by 412
Abstract
Cataracts are diseases characterized by the opacity of the ocular lens and the subsequent deterioration of vision. Metal ions are one of the factors that have been reported to induce crystallin aggregation. For HγS crystallin, several equivalent ratios of Cu(II) promote protein aggregation. [...] Read more.
Cataracts are diseases characterized by the opacity of the ocular lens and the subsequent deterioration of vision. Metal ions are one of the factors that have been reported to induce crystallin aggregation. For HγS crystallin, several equivalent ratios of Cu(II) promote protein aggregation. However, reports on zinc are contradictory. To characterize the process of metal ion binding and subsequent HγS crystallin aggregation, we performed dynamic light scattering, turbidimetry, isothermal titration calorimetry, fluorescence, and nuclear magnetic resonance experiments. The data show that both metal ions have multiple binding sites and promote aggregation. Zinc interacts mainly with the N-terminal domain, inducing small conformational changes, while copper interacts with both domains and induces unfolding, exposing the tryptophan residues to the solvent. Our work provides insight into the mechanisms of metal-induced aggregation at one of the lowest doses that appreciably promote aggregation over time. Full article
(This article belongs to the Section Biomacromolecules: Proteins)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HγS crystallin structure and sequence. The three-dimensional structure of HγS crystallin is shown (PDB id 2M3T), highlighting the N-terminal domain and the C-terminal domain. The sequence is shown at the bottom, with cysteines marked in yellow, histidines in blue, and tryptophans in magenta.</p>
Full article ">Figure 2
<p>Effect of metal ions on the aggregation of HγS crystallin. (<b>A</b>) DLS size distribution diagram by volume of HγS crystallin (black), HγS + Zn(II) (red), and HγS + Cu(II) (blue). A shift to the right can be observed, indicating the formation of large aggregates. (<b>B</b>) Turbidity assays of HγS crystallin in the absence (gray) or presence of Cu(II) (blue) and Zn(II) (red).</p>
Full article ">Figure 3
<p>Metal binding by ITC. (<b>A</b>) Isothermal titration calorimetry of HγS crystallin bound to Cu(II). (<b>B</b>) Isothermal titration calorimetry of HγS crystallin bound to Zn(II). The left side shows the experimental isothermal titrations, whereas the right side shows the reaction heat. Both metals exhibited complex behavior involving several processes accounting for the heat reactions. (<b>C</b>) DLS correlograms of HγS crystallin in the absence (gray) and presence of Cu(II) (blue) after ITC titration. A shift to the right (indicated with an arrow) can be observed, indicating the formation of large aggregates. (<b>D</b>) DLS correlograms of HγS crystallin in the absence (gray) and presence of Zn(II) (red) after ITC titration. A shift to the right (indicated with an arrow) can be observed, indicating the formation of large aggregates.</p>
Full article ">Figure 4
<p>Effect of Cu(II) and Zn(II) ions on the folding of HγS crystallin. (<b>A</b>) Fluorescence intensity ratio 350/325 as a function of time at 37 °C of HγS crystallin in the absence (gray), presence of Cu(II) (blue) and Zn(II) (red). (<b>B</b>) Normalized fluorescence spectra over time of HγS crystallin in the presence of 1.5 equivalent ratios of Cu(II). A shift to the right (indicated with an arrow) can be observed, indicating the exposure of tryptophan residues to the solvent. (<b>C</b>) Normalized fluorescence spectra over time of HγS crystallin in the presence of 1.5 equivalent ratios of Zn(II). First, a shift to the right can be observed, indicating the exposure of tryptophan residues to the solvent; then, a shift to the left indicates the burial of tryptophan residues. The direction of the shifts is indicated with arrows.</p>
Full article ">Figure 5
<p>Characterization of the HγS crystallin—Cu(II) interaction by NMR. (<b>A</b>) Overlay of a region of <sup>1</sup>H-<sup>15</sup>N HSQC spectrum of free HγS crystallin (black) and HγS crystallin in the presence of 0.5 (light blue) and 1.5 equivalent ratios of Cu(II) (blue). (<b>B</b>) Cu(II)-induced weighted intensity changes (I/I<sub>0</sub>) are mapped onto the HγS crystallin structure, color-coded via a linear ramp from cyan (no change) to blue (maximal signal loss). The residues for which the effect of copper binding could not be determined are shown in light gray. (<b>C</b>) Induced weighted intensity changes (I/I<sub>0</sub>) induced by the binding of Cu(II) per residue.</p>
Full article ">Figure 6
<p>Characterization of the HγS crystallin-Zn(II) interaction by NMR. (<b>A</b>) Overlay of a region of the <sup>1</sup>H-<sup>15</sup>N HSQC spectrum of free HγS crystallin (black) and HγS crystallin in the presence of 0.5 (light red) and 1.5 equivalent ratios of Zn(II) (red). (<b>B</b>) Zn(II)-induced chemical shift perturbations (CSPs) are mapped onto the HγS crystallin structure, color-coded via a linear ramp from cyan (no change) to red (maximal signal loss). The residues for which the effect of zinc-binding could not be determined are shown in light gray. (<b>C</b>) Chemical shift perturbations induced by the binding of Zn(II) per residue.</p>
Full article ">
12 pages, 3618 KiB  
Article
Integrating L-Cys-AuNCs in ZIF-8 with Enhanced Fluorescence and Strengthened Stability for Sensitive Detection of Copper Ions
by Ting Zhou, Luyao Zang, Xia Zhang, Xia Liu, Zijie Qu, Guodong Zhang, Xiufeng Wang, Fang Wang and Zhiqing Zhang
Molecules 2024, 29(24), 6011; https://doi.org/10.3390/molecules29246011 - 20 Dec 2024
Viewed by 435
Abstract
Gold nanoclusters (AuNCs) have been widely investigated because of their unique photoluminescence properties. However, the applications of AuNCs are limited by their poor stability and relatively low fluorescence. In the present work, we developed nanocomposites (L-Cys-AuNCs@ZIF-8) with high fluorescence and stability, which were [...] Read more.
Gold nanoclusters (AuNCs) have been widely investigated because of their unique photoluminescence properties. However, the applications of AuNCs are limited by their poor stability and relatively low fluorescence. In the present work, we developed nanocomposites (L-Cys-AuNCs@ZIF-8) with high fluorescence and stability, which were constructed by encapsulating the water-dispersible L-Cys-AuNCs into a ZIF-8 via Zn2+-triggered growth strategy without high temperature and pressure. The maximum emission wavelength of the L-Cys-AuNCs@ZIF-8 composite was at 868 nm, and the fluorescence intensity of L-Cys-AuNCs@ZIF-8 was nearly nine-fold compared with L-Cys-AuNCs without the ZIF-8 package. The mechanism investigation by fluorescence spectroscopy and X-ray photoelectron spectroscopy showed that L-Cys-AuNCs@ZIF-8 impeded ligand rotation, induced energy dissipation, and diminished the self-quenching effect, attributing to the spatial distribution of L-Cys-AuNCs. Based on the high fluorescence efficiency of L-Cys-AuNCs@ZIF-8, a “signal off” detective platform was proposed with copper ions as a model analyte, achieving a sensitive detection limit of Cu2+ at 16.7 nM. The quenching mechanism was confirmed, showing that the structure of the L-Cys-AuNCs@ZIF-8 nanocomposites was collapsed by the addition of Cu2+. Attributing to the strong adsorption ability between copper ions and pyridyl nitrogen, the as-prepared L-Cys-AuNCs@ZIF-8 was shown to accumulate Cu2+, and the Zn2+ in ZIF-8 was replaced by Cu2+. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration for preparation of L-Cys-AuNCs@ZIF-8 (<b>a</b>); TEM images of L-Cys-AuNCs (<b>b</b>), the structure of L-Cys-AuNCs which was precipitated by Zn<sup>2+</sup> (<b>c</b>) and L-Cys-AuNCs@ZIF-8 (<b>d</b>), respectively; high resolution transmission electron microscope image and the corresponding elemental (Zn, Au, S, N, O) mappings of L-Cys-AuNCs@ZIF-8 (<b>e</b>).</p>
Full article ">Figure 2
<p>Fluorescence intensity of L-Cys-AuNCs and L-Cys-AuNCs@ZIF-8 (<b>a</b>); UV-vis spectra of ZIF-8, L-Cys-AuNCs, and L-Cys-AuNCs@ZIF-8 (<b>b</b>); X-ray diffraction (XRD) patterns of L-Cys, L-Cys-AuNCs, ZIF-8, and L-Cys-AuNCs@ZIF-8 (<b>c</b>); Fourier transform infrared (FT-IR) spectra of L-Cys, L-Cys-AuNCs, ZIF-8, and L-Cys-AuNCs@ZIF-8 (<b>d</b>); XPS of L-Cys-AuNCs@ZIF-8 (<b>e</b>) and L-Cys-AuNCs (<b>f</b>), respectively.</p>
Full article ">Figure 3
<p>X-ray photoelectron spectra of Au (<b>a</b>), S (<b>b</b>), Zn (<b>c</b>), C (<b>d</b>), N (<b>e</b>), and O (<b>f</b>) in L-Cys-AuNCs@ZIF-8 and L-Cys-AuNCs, respectively.</p>
Full article ">Figure 4
<p>Zeta potential values of L-Cys-AuNCs, ZIF-8, and L-Cys-AuNCs@ZIF-8 (<b>a</b>); stability of L-Cys-AuNCs and L-Cys-AuNCs@ZIF-8 at different temperatures (<b>b</b>), after long-term storage (<b>c</b>), and at different concentrations of NaCl (<b>d</b>).</p>
Full article ">Figure 5
<p>Fluorescence intensity evolution of L-Cys-AuNCs@ZIF-8 in the presence of different amounts of Cu<sup>2+</sup> (<b>a</b>); <span class="html-italic">F</span>/<span class="html-italic">F</span><sub>0</sub> versus concentrations of Cu<sup>2+</sup> (<b>b</b>); the linear relationship between <span class="html-italic">F</span>/<span class="html-italic">F</span><sub>0</sub> and Cu<sup>2+</sup> at low concentrations (<b>c</b>); interference study with different metal ions (<b>d</b>).</p>
Full article ">Figure 6
<p>Fluorescence intensity and photograph of L-Cys-AuNCs@ZIF-8 and L-Cys-AuNCs@ZIF-8 + Cu<sup>2+</sup>, respectively (<b>a</b>); TEM image of L-Cys-AuNCs@ZIF-8 mixed with Cu<sup>2+</sup> (<b>b</b>); XRD spectrum (<b>c</b>) and FT-IR spectrum (<b>d</b>) of L-Cys-AuNCs@ZIF-8 + Cu<sup>2+</sup>.</p>
Full article ">Scheme 1
<p>(<b>A</b>) Synthesis of L-Cys-AuNCs; (<b>B</b>) illustration showing the fabrication of L-Cys-AuNCs@ZIF-8 with improved fluorescence; (<b>C</b>) the as-prepared L-Cys-AuNCs@ZIF-8, used for Cu<sup>2+</sup> detection.</p>
Full article ">
Back to TopTop