Nothing Special   »   [go: up one dir, main page]

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (464)

Search Parameters:
Keywords = DOT1L

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 6161 KiB  
Article
Screening of a Fraction with Higher Amyloid β Aggregation Inhibitory Activity from a Library Containing 210 Mushroom Extracts Using a Microliter-Scale High-Throughput Screening System with Quantum Dot Imaging
by Gegentuya Huanood, Mahadeva M. M. Swamy, Rina Sasaki, Keiya Shimamori, Masahiro Kuragano, Enkhmaa Enkhbat, Yoshiko Suga, Masaki Anetai, Kenji Monde and Kiyotaka Tokuraku
Foods 2024, 13(23), 3740; https://doi.org/10.3390/foods13233740 - 22 Nov 2024
Viewed by 375
Abstract
Alzheimer’s disease (AD) is a highly prevalent neurodegenerative disease hallmarked by amyloid plaques and neurofibrillary tangles. Amyloid plaques are formed by the amyloid β (Aβ) aggregation, so substances that inhibit this aggregation are useful for preventing and treating AD. Mushrooms are widely used [...] Read more.
Alzheimer’s disease (AD) is a highly prevalent neurodegenerative disease hallmarked by amyloid plaques and neurofibrillary tangles. Amyloid plaques are formed by the amyloid β (Aβ) aggregation, so substances that inhibit this aggregation are useful for preventing and treating AD. Mushrooms are widely used medicinal fungi with high edible and nutritional value. Mushrooms have a variety of biologically active ingredients, and studies have shown that they have certain effects in anti-bacterial, anti-oxidation, anti-inflammatory, anti-tumor, and immune regulation. Previously, we developed a microliter-scale high-throughput screening (MSHTS) system using quantum dot (QD) nanoprobes to screen Aβ aggregation inhibitors. In this study, we appraised the Aβ aggregation inhibitory activity of 210 natural mushrooms from Hokkaido (Japan) and found 11 samples with high activity. We then selected Elfvingia applanata and Fuscoporia obliqua for extraction and purification as these samples were able to suppress Aβ-induced neurocytotoxicity and were readily available in large quantities. We found that the ethyl acetate (EtOAc) extract of E. applanata has high Aβ aggregation inhibitory activity, so we performed silica gel column chromatography fractionation and found that fraction 5 (f5) of the EtOAc extract displayed the highest Aβ aggregation inhibitory activity among all mushroom samples. The half-maximal effective concentration (EC50) value was 2.30 µg/mL, higher than the EC50 of 10.7 µg/mL for rosmarinic acid, a well-known Aβ aggregation inhibitor. This inhibitory activity decreased with further purification, suggesting that some compounds act synergistically. The f5 fraction also inhibited the deposition of Aβ aggregates on the cell surface of human neuroblastoma SH-SY5Y cells. Our expectation is that f5, with additional tests, may eventually prove to be an inhibitor for the prevention of AD. Full article
Show Figures

Figure 1

Figure 1
<p>Fluorescence images and EC<sub>50</sub> values demonstrating the inhibitory effects of mushroom extracts on Aβ aggregation. (<b>A</b>) Analysis of Aβ aggregation inhibitory activity of 11 mushroom extracts. Mixed solutions of 25 μM Aβ<sub>42</sub> and mushroom extracts were added to 1536-well plates and imaged by fluorescence microscopy after incubation at 37 °C for 24 h. The captured images were cropped to 432 × 432 pixels. (<b>B</b>) Measurement of EC<sub>50</sub> values of 11 mushroom extracts. Data are represented as mean ± SD (<span class="html-italic">n</span> = 4). Smaller EC<sub>50</sub> values indicate higher inhibitory activity. EC<sub>50</sub> values were calculated using Prism GraphPad software.</p>
Full article ">Figure 2
<p>Inhibitory effects of 11 mushroom extracts on Aβ-induced PC12 cell toxicity. PC12 cells were differentiated using NGF for 24 h and treated with 25 μM Aβ<sub>42</sub> containing each mushroom extract for 24 h. Cell viability measured by the MTT assay is shown as the relative percentage of absorbance of treated samples compared to the control without Aβ and mushroom extract. Each extract was compared with Aβ<sub>42</sub>. Each plot and bar graph represents the mean ± SD (<span class="html-italic">n</span> = 3 separate experiments with extracts, six separate experiments in 25 μM Aβ only). (*: <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, Welch’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 3
<p>Fluorescence images and EC<sub>50</sub> values of the inhibitory effect of extracts of 2 mushrooms with different solvents on Aβ aggregation. (<b>A</b>) Analysis of the inhibitory activity of t037 at different concentrations on Aβ aggregation in five extraction solvents. (<b>B</b>) Analysis of the inhibitory activity of t100 at different concentrations on Aβ aggregation in five extraction solvents. The fluorescence images were cropped to 432 × 432 pixels. (<b>C</b>) Measurement of EC<sub>50</sub> values of t037 and t100 in five extraction solvents. Samples with SD values not less than 50% were considered as unable to calculate EC<sub>50</sub> values and expressed as not determined (ND). Data are represented as the mean ± SD (<span class="html-italic">n</span> = 3).</p>
Full article ">Figure 4
<p>Evaluation of the inhibitory effects of active ingredients in the t037 EtOAc extract. (<b>A</b>) The t037 EtOAc extract was separated into five fractions (f1−f5) by silica gel column fractionation. f5 was further separated into four fractions (f5f1−f5f4) by same silica gel column fractionation. (<b>B</b>) Fluorescence imaging of the two fractions (f5 and f5f3) with high Aβ aggregation inhibitory activity. The fluorescence images were cropped to 432 × 432 pixels. (<b>C</b>) The inhibition curve of f5. (<b>D</b>) The inhibition curve of f5f3. The inhibition curves were drawn from the SD values, and EC<sub>50</sub> was calculated using Prism GraphPad software.</p>
Full article ">Figure 5
<p>Inhibitory effect of the EtOAc extract of t037 on extracellular Aβ deposition in SH-SY5Y cells. (<b>A</b>) SH-SY5Y cells were co-cultured with 25 μM Aβ and 25 nM QDAβ and different concentrations of f5 and f5f3. Images were collected using an inverted fluorescence microscope. 1.2% DMSO was used as the negative control (without Aβ). DIA: bright field image; QD: fluorescence image of QD and Aβ aggregation; Merge: merged image. (<b>B</b>) The QD field’s mean gray value of each group was determined. Data are represented as mean ± SD (<span class="html-italic">n</span> = 3). (**: <span class="html-italic">p</span> &lt; 0.01, ***: <span class="html-italic">p</span> &lt; 0.001, Welch’s <span class="html-italic">t</span>-test).</p>
Full article ">
17 pages, 6937 KiB  
Article
Characterization of Particle Size Effects on Sintering Shrinkage and Porosity in Stainless Steel Metal Injection Molding Using Multi-Physics Simulation
by Ying Wu, Kaibo Guo and Junfang Ni
Materials 2024, 17(23), 5691; https://doi.org/10.3390/ma17235691 - 21 Nov 2024
Viewed by 270
Abstract
In this study, three stainless steel materials (17-4PH, 316L, and 304) were experimentally simulated using metal injection molding (MIM) technology to explore the size shrinkage behavior and defect formation mechanism of materials with different particle sizes during sintering. The sintering environment was linearly [...] Read more.
In this study, three stainless steel materials (17-4PH, 316L, and 304) were experimentally simulated using metal injection molding (MIM) technology to explore the size shrinkage behavior and defect formation mechanism of materials with different particle sizes during sintering. The sintering environment was linearly heated to 1250 °C at a rate of 5 °C/min and kept warm for 90 min. Multi-physics field coupling analysis was performed using ANSYS Workbench software. Two different regions were selected to simulate the total deformation trend of the material during sintering. The simulation results were compared with data from SEM and EDS analyses to elucidate the influence of particle size on shrinkage behavior and defect distribution. The findings indicate that the gaps between particles far away from the gate position became larger, the degree of densification decreased, the porosity was higher, and the number of white dot inclusions increased. Among the three materials, 17-4PH, which had the smallest particle size, had a greater sintering driving force, a better degree of densification, a smaller predicted total deformation, and a higher shrinkage rate, which is consistent with the hardness test data and the actual density data. In addition, the densification advantage of small particle size powder is not only related to surface energy but is also closely linked to the uniformity of its microstructure. The analysis in this study further promotes the performance optimization of stainless steel materials, indicates a scientific basis for future process improvements and high-precision parts manufacturing in MIM technology, and points to the development direction for high-performance materials. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the (<b>a</b>) 17-4PH powder, (<b>b</b>) 316L powder, and (<b>c</b>) 304 powder.</p>
Full article ">Figure 2
<p>The fixed support surface: gate (<b>a</b>) and ANSYS model (<b>b</b>).</p>
Full article ">Figure 3
<p>A magnified view of the 17-4PH results: total deformation side (<b>a</b>) and gate side (<b>b</b>).</p>
Full article ">Figure 4
<p>SEM images of 17-4PH’s near gate and far gate areas. The arrows point to the pores observed in the material.</p>
Full article ">Figure 5
<p>Magnified results of the total deformation of 306L (<b>a</b>) and 304 (<b>b</b>).</p>
Full article ">Figure 6
<p>SEM images of 316L (<b>a</b>) and 304 (<b>b</b>) materials’ near and far gates. The arrows indicate the pores observed in the material.</p>
Full article ">Figure 7
<p>Comparison of actual and theoretical densities of 17-4PH, 316L, and 304.</p>
Full article ">Figure 8
<p>EDS spectrum analysis of 17-4PH (<b>a</b>), 316L (<b>b</b>), and 304 (<b>c</b>).</p>
Full article ">Figure 8 Cont.
<p>EDS spectrum analysis of 17-4PH (<b>a</b>), 316L (<b>b</b>), and 304 (<b>c</b>).</p>
Full article ">Figure 9
<p>Dimensional shrinkage and standard deviation of 17-4PH, 316L, and 304 stainless steel in the X, Y, and Z directions.</p>
Full article ">Figure 10
<p>The Vickers hardness analysis results and the standard deviation from the near gate to the far gate.</p>
Full article ">
13 pages, 3147 KiB  
Article
Photocatalytic Degradation of Methylene Blue by Surface-Modified SnO2/Se-Doped QDs
by Luis Alamo-Nole and Sonia J. Bailon-Ruiz
Micro 2024, 4(4), 721-733; https://doi.org/10.3390/micro4040044 - 21 Nov 2024
Viewed by 246
Abstract
Developing new nanomaterials and performing functionalization to increase their photocatalytic capacity are essential in developing low-cost, eco-friendly, and multipurpose-capacity catalysts. In this research, SnO2/Se-doped quantum dots (QDs) covered with glycerol (SnO2/Se-GLY) were synthesized using microwave irradiation. Then, their cover [...] Read more.
Developing new nanomaterials and performing functionalization to increase their photocatalytic capacity are essential in developing low-cost, eco-friendly, and multipurpose-capacity catalysts. In this research, SnO2/Se-doped quantum dots (QDs) covered with glycerol (SnO2/Se-GLY) were synthesized using microwave irradiation. Then, their cover was replaced with glutaraldehyde through a ligand exchange procedure (SnO2/Se-GLUT). The XRD analyses confirmed a tetragonal rutile structure of SnO2. The HR-TEM analysis confirmed the generation of QDs with a size around 8 nm, and the optical analysis evidenced low bandgap energies of 3.25 and 3.26 eV for the SnO2/Se-GLY and SnO2/Se-GLUT QDs, respectively. Zeta-sizer analysis showed that the hydrodynamic sizes for both nanoparticles were around 230 nm (50 mg/L), and the zeta potential confirmed that SnO2/Se-GLUT QDs were more stable than SnO2/Se-GLY QDs. The cover-modified QDs (SnO2/Se-GLUT) showed a higher and faster adsorption capacity, followed by a slower photocatalytic process than the original QDs (SnO2/Se-GLY). The QTOF-LC-MS analysis confirmed MB degradation through the identification of intermediates such as azure A, azure B, azure C, and phenothiazine. Adsorption isotherm analysis indicated Langmuir model compliance, supporting the high monolayer adsorption capacity and efficiency of these QDs as adsorbent/photocatalytic agents for organic pollutant removal. This dual capability for adsorption and photodegradation, along with the demonstrated reusability, highlights the potential of SnO2/Se QDs in wastewater treatment and environmental remediation. Full article
(This article belongs to the Special Issue Advances in Micro- and Nanomaterials: Synthesis and Applications)
Show Figures

Figure 1

Figure 1
<p>SnO<sub>2</sub>/Se-GLY (<b>left</b>) and SnO<sub>2</sub>/Se-GLUT (<b>right</b>) nanoparticles.</p>
Full article ">Figure 2
<p>The absorption spectrum of the SnO<sub>2</sub>/Se-GLY nanoparticles. Inset: plot of (<span class="html-italic">αhν</span>)<sup>2</sup> vs. (<span class="html-italic">hν</span>) for the direct transition.</p>
Full article ">Figure 3
<p>The absorption spectrum of the SnSe-GLUT nanoparticles. Inset: plot of (<span class="html-italic">αhν</span>)<sup>2</sup> vs. (<span class="html-italic">hν</span>) for the direct transition.</p>
Full article ">Figure 4
<p>FT-IR spectra of pure glycerol, glutaraldehyde, and cover residues on the SnO<sub>2</sub>/Se-GLY and SnO<sub>2</sub>/Se-GLUT QDs.</p>
Full article ">Figure 5
<p>Powder X-ray diffraction pattern of SnO<sub>2</sub>/Se-doped QDs covered with glycerol. The orange lines are the standard lines for the rutile structure cassiterite phase.</p>
Full article ">Figure 6
<p>TEM image (<b>a</b>) and electron diffraction pattern (<b>b</b>) of the SnO<sub>2</sub>/Se-GLY QDs. Dashed circles indicate the particles.</p>
Full article ">Figure 7
<p>Hydrodynamic size of the SnO<sub>2</sub>/Se-GLY (<b>a</b>) and SnO<sub>2</sub>/Se-GLUT QDs (<b>b</b>). Attenuator at 9 and cuvette position at 4.64 mm.</p>
Full article ">Figure 8
<p>Adsorption and photodegradation of low methylene blue concentration using SnO<sub>2</sub>/Se-GLY and SnO<sub>2</sub>/Se-GLUT QDs.</p>
Full article ">Figure 9
<p>Adsorption and photodegradation of high methylene blue concentration using SnO<sub>2</sub>/Se-GLY and SnO<sub>2</sub>/Se-GLUT QDs.</p>
Full article ">Figure 10
<p>Adsorption isotherm (<b>a</b>) and Langmuir isotherm model (<b>b</b>) of methylene blue adsorption on the SnO<sub>2</sub>/Se-GLUT QDs.</p>
Full article ">Figure 11
<p>Recycling capacity of SnO<sub>2</sub>/Se-GLUT QDs for four photodegradation cycles of methylene blue at 5 µM concentration.</p>
Full article ">
13 pages, 2931 KiB  
Article
Ocular Mucous Membrane Pemphigoid Demonstrates a Distinct Autoantibody Profile from Those of Other Autoimmune Blistering Diseases: A Preliminary Study
by Yingzi Liu, Lei Bao, Dharm Sodha, Jing Li, Adrian Mansini, Ali R. Djalilian, Xiaoguang Li, Hua Qian, Norito Ishii, Takashi Hashimoto and Kyle T. Amber
Antibodies 2024, 13(4), 91; https://doi.org/10.3390/antib13040091 - 14 Nov 2024
Viewed by 408
Abstract
Background: Ocular predominant mucous membrane pemphigoid (oMMP) is a severe subtype of autoimmune blistering disease (AIBD), which can result in scarring and vision loss. The diagnosis of oMMP is challenging as patients often have undetectable levels of circulating autoantibodies by conventional assays. [...] Read more.
Background: Ocular predominant mucous membrane pemphigoid (oMMP) is a severe subtype of autoimmune blistering disease (AIBD), which can result in scarring and vision loss. The diagnosis of oMMP is challenging as patients often have undetectable levels of circulating autoantibodies by conventional assays. Likewise, the principal autoantigen in oMMP has been an area of debate. Methods: In this preliminary experiment, we performed Phage Immunoprecipitation Sequencing (PhIP-seq) on sera from patients with oMMP, as well as non-ocular MMP, bullous pemphigoid, and mucocutaneous-type pemphigus vulgaris. Results: We identified several autoantigens unique to oMMP relative to other AIBDs. We then cross-referenced these antigens against previously published single-nuclei datasets, as well as the International Mouse Phenotyping Consortium Database. Several protein hits identified in our study demonstrated enriched expression on the anterior surface epithelia, including TNKS1BP1, SEC16B, FNBP4, CASZ1, GOLGB1, DOT1L, PRDM 15, LARP4B, and RPL6. Likewise, a previous study of mouse knockout models of murine analogs CASZ1, HIP1, and ELOA2 reported that these mice showed abnormalities in terms of the ocular surface and development in the eyes. Notably, PhIP-seq failed to identify the canonical markers of AIBDs such as BP180, BP230, desmogleins 1 and 3, or integrin β4, indicating that the patient autoantibodies react with conformational epitopes rather than linear epitopes. Conclusions: oMMP patients demonstrate a unique autoantibody repertoire relative to the other AIBDs. Further validation of the identified autoantibodies will shed light on their potentially pathogenic role. Full article
(This article belongs to the Section Humoral Immunity)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Quality control of PhIP-Seq data. (<b>A</b>) Number of mapped counts for each sample. The red dashed line indicates the level of 5× mean coverage. (<b>B</b>) Number of mapped counts for each group. (<b>C</b>) Number of peptide EdgeR hits called in each sample. (<b>D</b>) Number of peptide EdgeR hits called in each group. (<b>E</b>) Heatmap of pipeline-called anti-GFAP (glial fibrillary acidic protein) positive control hits in all samples.</p>
Full article ">Figure 2
<p>Distinct autoantibodies in various AIBD sera. (<b>A</b>) Heatmap of distinct autoantibodies defined from Fisher exact test on protein hits in the group of interest vs. controls. (<b>B</b>) Dot plot showing the expression of the top twenty distinct autoantibodies in MMP compared to other AIBD diseases in corneoscleral wedge (CSW) cells. Cell types were annotated at the top and color-coded the same as in panel (<b>C</b>). (<b>C</b>) UMAP visualization of clustering of nuclei derived from CSW tissue (top left). Cell type and cell numbers were shown (bottom left). Feature plots showing expression of selected top gene identified in MMP compared to other AIBD diseases in all CSW cells (right). Ocular surface epithelium subtypes are outlined.</p>
Full article ">Figure 3
<p>GO analysis of top autoantibodies enriched in oMMP patient sera. Bubble plots showing the enriched terms of biological process (<b>left</b>) and cellular components (<b>right</b>) in MMP compared to other AIBD diseases. Adjusted <span class="html-italic">p</span>-value is indicated in y-axis. Size of the bubbles indicates the number of genes enriched in each term.</p>
Full article ">Figure 4
<p>Autoantibodies enriched in mcPV patient sera. (<b>A</b>) Venn diagram representing the numbers and overlapping of autoantibodies identified from PhIP-seq in the current study and the study by Kalantari-Dehaghi et al. [<a href="#B29-antibodies-13-00091" class="html-bibr">29</a>]; (<b>B</b>) Table of four overlapping autoantibodies between the two studies and proportion of patients showing autoantibodies in disease and control groups.</p>
Full article ">
18 pages, 2199 KiB  
Article
A New Chemiluminescence Assay for Hypochlorite Detection in Water: A Synergistic Combination of WS2 Quantum Dots and Luminol
by Madina M. Sozarukova, Elena V. Proskurnina, Ekaterina M. Kochneva, Andrey K. Barinov, Alexander E. Baranchikov and Vladimir K. Ivanov
Water 2024, 16(21), 3044; https://doi.org/10.3390/w16213044 - 24 Oct 2024
Viewed by 569
Abstract
The issue of the qualitative and quantitative analysis of the concentration of oxidising species in aquatic environments is crucial for a wide range of biological and environmental tasks. In particular, reactive chlorine species, specifically hypochlorite (ClO), play a significant biochemical role [...] Read more.
The issue of the qualitative and quantitative analysis of the concentration of oxidising species in aquatic environments is crucial for a wide range of biological and environmental tasks. In particular, reactive chlorine species, specifically hypochlorite (ClO), play a significant biochemical role in the operation of the immune system. There is also the challenge of determining the presence of ClO in purified drinking water that is supplied by water treatment systems. Traditional chemical analytical methods often lack the required selectivity and sensitivity to detect oxidising compounds, and chemiluminescence-based techniques offer an alternative solution. In this study, we propose a simple and selective approach for the chemiluminescent detection of hypochlorite in aqueous media under neutral conditions. The technique is based on measuring a chemiluminescent signal generated in the presence of hypochlorite by a combined probe comprising commercially available WS2 quantum dots and luminol. The oxidation of WS2 with hypochlorite followed by a reaction with luminol results in an intense luminescent signal that enables the selective determination of hypochlorite under neutral conditions. The greatest sensitivity with this method was achieved when combining WS2 quantum dots with L-012, a highly sensitive analogue of luminol. Additionally, the use of L-012 improved the detection limit for hypochlorite to 2 × 10−6 M. Due to its selectivity in determining hypochlorite in the presence of reactive oxygen species (hydrogen peroxide) under neutral conditions with high sensitivity and with a wide linear range, the proposed approach provides an attractive analytical tool for the analysis of water samples and biological liquids. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) UV absorption and (<b>b</b>) fluorescence spectra (λ<sub>ex</sub> = 350 nm) of WS<sub>2</sub> QDs.</p>
Full article ">Figure 2
<p>Chemiluminescent kinetic curves for the NaClO solution (1.5 mM) after the addition of different probes: (<b>a</b>) luminol (60 µM); (<b>b</b>) WS<sub>2</sub> QDs (60 µM); and (<b>c</b>) luminol (30 µM) and WS<sub>2</sub> QDs (60 µM). The aliquot volume of each component is 20 µL. In the figures, red arrows indicate the injection of the luminol solution.</p>
Full article ">Figure 3
<p>(<b>a</b>) Typical chemiluminescent signal of the WS<sub>2</sub> QDs aqueous colloid solution (20 µL, 60 µM) upon the sequential addition of the sodium hypochlorite solution (20 µL, 1.5 mM) and luminol (20 µL, 30 µM)—the time of interaction between WS<sub>2</sub> QDs and hypochlorite was 2 min. (<b>b</b>) The areas under the chemiluminescence curves (lightsums) for various times of interaction between NaClO with WS<sub>2</sub> QDs before the addition of luminol.</p>
Full article ">Figure 4
<p>Chemiluminescent kinetic curves for the NaClO (20 µL, 1.5 mM) solution upon the addition of (<b>a</b>) WS<sub>2</sub> QDs and (<b>b</b>) luminol at various concentrations. The areas under the corresponding chemiluminescence curves (lightsums) are as a function of the concentration of (<b>c</b>) WS<sub>2</sub> QDs and (<b>d</b>) luminol.</p>
Full article ">Figure 5
<p>(<b>a</b>) Chemiluminescent kinetic curves for the NaClO (20 µL, 1.5 mM) solution upon the addition of WS<sub>2</sub> QDs and luminol with WS<sub>2</sub> QDs-to-luminol molar ratios of 1:1, 1:2, and 2:1; (<b>b</b>) the areas under the corresponding chemiluminescence curves (lightsums).</p>
Full article ">Figure 6
<p>(<b>a</b>) Chemiluminescence kinetic curves for the NaClO solution (20 µL, 1.5 mM) upon the addition of the WS<sub>2</sub> QDs (20 µL, 40 µM) /luminol (20 µL, 20 µM) probe followed by the addition of various quantities of sodium ascorbate; (<b>b</b>) chemiluminescence suppression (S<sub>CL</sub>, a.u.) on sodium ascorbate concentration.</p>
Full article ">Figure 7
<p>(<b>a</b>) Chemiluminescent kinetic curves for the solutions of NaClO with various concentrations; (<b>b</b>) the area under the corresponding chemiluminescence curve (lightsums, S) as a function of hypochlorite concentration; (<b>c</b>) the linear relationship between the analytical signal (lightsum) and the concentration of hypochlorite in the range of 0.25–1.80 mM.</p>
Full article ">Figure 8
<p>(<b>a</b>) Chemiluminescent kinetic curves for solutions of NaClO with various concentrations; (<b>b</b>) the linear relationship between the analytical signal (lightsum) and the concentration of hypochlorite in the range of 2.5–150 µM.</p>
Full article ">Figure 9
<p>(<b>a</b>) A typical chemiluminescent kinetic curve for the solution of H<sub>2</sub>O<sub>2</sub> (1.5 mM) in the presence of luminol (20 µM); (<b>b</b>) histogram of the lightsum values for H<sub>2</sub>O<sub>2</sub> solutions after the addition of different chemiluminescent probes, namely luminol (20 µM), WS<sub>2</sub> QDs (40 µM), and luminol (20 µM)/WS<sub>2</sub> QDs (40 µM). The aliquot volume of each component is 20 µL.</p>
Full article ">Figure 10
<p>A typical chemiluminescent kinetic curve for the solution of H<sub>2</sub>O<sub>2</sub> (1.5 mM)/Fe(II) (100 µM) upon the addition of luminol (20 µM); WS<sub>2</sub> QDs (40 µM); and luminol (20 µM) and WS<sub>2</sub> QDs (40 µM). The aliquot volume of each component is 20 µL.</p>
Full article ">
15 pages, 5855 KiB  
Article
Electrospun Nanofiber Dopped with TiO2 and Carbon Quantum Dots for the Photocatalytic Degradation of Antibiotics
by Valentina Silva, Diana L. D. Lima, Etelvina de Matos Gomes, Bernardo Almeida, Vânia Calisto, Rosa M. F. Baptista and Goreti Pereira
Polymers 2024, 16(21), 2960; https://doi.org/10.3390/polym16212960 - 22 Oct 2024
Viewed by 587
Abstract
Novel photocatalysts were synthesized through the association of carbon quantum dots (CQDs) with commercial (P25) titanium dioxide (TiO2) by sonication. The resulting TiO2/CQDs composite was then incorporated into the polyamide 66 (PA66) biopolymer nanofibers using the electrospinning technique, considering [...] Read more.
Novel photocatalysts were synthesized through the association of carbon quantum dots (CQDs) with commercial (P25) titanium dioxide (TiO2) by sonication. The resulting TiO2/CQDs composite was then incorporated into the polyamide 66 (PA66) biopolymer nanofibers using the electrospinning technique, considering a composite nanoparticles-to-polymer ratio of 1:2 in the electrospinning precursor solution. The produced nanofibers presented suitable morphology and were tested for the photocatalytic degradation under simulated solar radiation of 10 mg L−1 of amoxicillin (AMX) and sulfadiazine (SDZ), in phosphate buffer solution (pH 8.06) and river water, using 1.5 g L−1 of photocatalyst. The presence of the photocatalyst increased the removal of AMX in phosphate buffer solution by 30 times, reducing the AMX degradation half-life time from 62 ± 1 h (without catalyst) to 1.98 ± 0.06 h. Moreover, SDZ degradation half-life time in phosphate buffer solution was reduced from 5.4 ± 0.1 h (without catalyst) to 1.87 ± 0.05 h in the presence of the photocatalyst. Furthermore, the PA66/TiO2/CQDs were also efficient in river water samples and maintained their performance in at least three cycles of SDZ photodegradation in river water. The presented results evidence that the produced photocatalyst can be a promising and sustainable solution for antibiotics’ efficient removal from water. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) System used for the production of nanofibers (E-fiber EF 100). (<b>B</b>) Photocatalytic experiment of AMX solution in PBS containing the PA66/TiO<sub>2</sub>/CQDs nanofibers.</p>
Full article ">Figure 2
<p>SEM images of PA66 (<b>A</b>,<b>B</b>) and PA66/TiO<sub>2</sub>/CQDs (<b>C</b>,<b>D</b>) nanofibers.</p>
Full article ">Figure 3
<p>EDS mapping of PA66/TiO<sub>2</sub>/CQDs nanofibers, confirming the presence of C and Ti.</p>
Full article ">Figure 4
<p>X-Ray diffraction pattern of TiO<sub>2</sub>, TiO<sub>2</sub>/CQDs and PA66/TiO<sub>2</sub>/CQDs composites.</p>
Full article ">Figure 5
<p>FTIR-ATR of PA66 and PA66/TiO<sub>2</sub>/CQDs nanofibers.</p>
Full article ">Figure 6
<p>Photodegradation kinetic experimental data together with pseudo-first-order fittings for the photodegradation of (<b>A</b>) AMX and (<b>B</b>) SDZ (10 mg L<sup>−1</sup>), in the absence of the photocatalyst (photolysis), in the presence of PA66 fibers (1.5 g L<sup>−1</sup>), and with the photocatalyst (PA66/TiO<sub>2</sub>/CQDs, 1.5 g L<sup>−1</sup>), in PBS pH 8.06 or river water. The error bar represents the standard deviation (n = 3).</p>
Full article ">Figure 7
<p>Reuse cycles of PA66/TiO<sub>2</sub>/CQDs (1.5 g L<sup>−1</sup>) in the photodegradation of SDZ (10 mg L<sup>−1</sup>), in the absence of light (Dark, grey bars) and with irradiation for 4 h (Light, blue bars), in river water. The error bar represents the standard deviation (<span class="html-italic">n</span> = 3) with one-way ANOVA analysis (** <span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
19 pages, 5242 KiB  
Article
Unlocking the Luminescent Potential of Fish-Scale-Derived Carbon Nanoparticles for Multicolor Conversion
by Najeeb S. Abdulla II, Marvin Jose F. Fernandez, Bakhytzhan Baptayev and Mannix P. Balanay
Int. J. Mol. Sci. 2024, 25(20), 10929; https://doi.org/10.3390/ijms252010929 - 11 Oct 2024
Viewed by 735
Abstract
This study introduces a novel approach to addressing environmental issues by developing fish-scale carbon nanoparticles (FSCNPs) with a wide range of colors from discarded fish scales. The process involves hydrothermally synthesizing raw tamban (Sardinella) fish scales sourced from Universal Canning, Inc. in Zamboanga [...] Read more.
This study introduces a novel approach to addressing environmental issues by developing fish-scale carbon nanoparticles (FSCNPs) with a wide range of colors from discarded fish scales. The process involves hydrothermally synthesizing raw tamban (Sardinella) fish scales sourced from Universal Canning, Inc. in Zamboanga City, Philippines. The optimization of the synthesis was achieved using the response surface methodology with a Box–Behnken design. The resulting FSCNPs exhibited unique structural and chemical properties akin to carbonized polymer dots, enhancing their versatility. The solid-state fluorescence of these nanoparticles can be modulated by varying their concentration in a polyvinylpyrrolidone matrix, yielding colors such as blue, green, yellow, and red-orange with Commission Internationale de l’Eclairage coordinates of (0.23, 0.38), (0.32, 0.43), (0.37, 0.43), and (0.46, 0.48), respectively. An analysis of the luminescence mechanism highlights cross-linking emissions, aggregation-induced emissions, and non-covalent interactions, which contribute to concentration-dependent fluorescence and tunable emission colors. These optical characteristics suggest that FSCNPs have significant potential for diverse applications, particularly in opto-electronic devices. Full article
(This article belongs to the Special Issue Properties and Applications of Nanoparticles and Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>–<b>c</b>) TEM images of FSCNPs and (<b>d</b>) particle size distribution statistics of the individual carbon dots.</p>
Full article ">Figure 2
<p>(<b>a</b>) XPS survey scan and high resolution XPS spectra of FSCNPs for (<b>b</b>) C1s, (<b>c</b>) N1s, and (<b>d</b>) O1s.</p>
Full article ">Figure 3
<p>(<b>a</b>) FTIR spectra of the FSCNP and raw fish scales. Deconvolution of FTIR spectra in (<b>b</b>) 4000–2600 cm<sup>−1</sup> and (<b>c</b>) 1800–1500 cm<sup>−1</sup> regions.</p>
Full article ">Figure 4
<p>(<b>a</b>) XRD and (<b>b</b>) Raman spectra of FSCNPs. The spectra also include the deconvoluted peaks for further identification.</p>
Full article ">Figure 5
<p>(<b>a</b>) The absorption and fluorescence (excitation and emission) spectra of FSCNPs. The inset shows the images when FSCNP is irradiated with regular light (left) and 365 nm light source (left). (<b>b</b>) Excitation-independent emission is observed at excitation wavelengths of 300–320 nm. (<b>c</b>) Excitation-dependent red-shift emission is observed when the excitation wavelength is varied from 330 to 400 nm.</p>
Full article ">Figure 6
<p>CIE 1931 diagram illustrating the chromaticity of the prepared LED devices. Images (<b>a</b>–<b>d</b>) show actual photographs of the LED light emissions irradiated at 365 nm, with each image plotted on the CIE diagram along with its corresponding x, y coordinates.</p>
Full article ">Figure 7
<p>Mechanism of formation and luminescence origin of FSCNPs.</p>
Full article ">Figure 8
<p>Concentration-dependent fluorescence of FSCNPs: Red-shifted emission and multicolor luminescence from 430 to 635 nm.</p>
Full article ">
12 pages, 3910 KiB  
Article
Fast and Sensitive Determination of Iodide Based on Ternary Chalcogenides Nanoparticles
by Zhitai Wang, Nengtao Wu, Weihao Wang, Yaozheng Hu, Zhijie Luo, Yuhui Zheng and Qianming Wang
Molecules 2024, 29(19), 4751; https://doi.org/10.3390/molecules29194751 - 8 Oct 2024
Viewed by 519
Abstract
A fluorescent probe based on ternary AgFeS2 quantum dots has been prepared for the design of ternary chalcogenides. The nanoparticles are synthesized with oleylamine as a stabilizer at a low temperature (particle size in the range of 2 to 3 nm) and [...] Read more.
A fluorescent probe based on ternary AgFeS2 quantum dots has been prepared for the design of ternary chalcogenides. The nanoparticles are synthesized with oleylamine as a stabilizer at a low temperature (particle size in the range of 2 to 3 nm) and they exhibit an intense blue emission in aqueous media. As for their internal structure, each nanoparticle’s relative stoichiometric ratio (AgFe1.01S1.91) is very close to the theoretical value of 1:1:2. Their magnetic properties have been studied with a vibrating sample magnetometer and they have ferromagnetism between 4 K and 298 K (applied magnetic field ranging between −10,000 and 10,000 Oe). In the presence of iodide ions, the emission at 458 nm derived from AgFeS2 QDs has been observed to give rise to fluorescence quenching. The detection system is based on a static quenching process and morphological change between iodide ions and AgFeS2, which has a good linear range from 0 to 37.5 μmol/L, with a limit of detection of 0.99 μM. The nanoprobe responds within 30 s for the efficient detection of iodide. Such functional quantum dots will provide a powerful indicator in environmental and bio-sensing applications. Full article
Show Figures

Figure 1

Figure 1
<p>The X-ray diffraction (XRD) patterns of the obtained samples; the blue vertical is the standard literature data.</p>
Full article ">Figure 2
<p>The EDX spectrum of AgFeS<sub>2</sub> QDs (the bottom table represents the elemental composition of AgFeS<sub>2</sub>).</p>
Full article ">Figure 3
<p>(<b>a</b>) HR-TEM images of the AgFeS<sub>2</sub> QDs and (<b>b</b>) the DLS curve of the AgFeS<sub>2</sub> QDs. The inset in the top right corner is the corresponding high-resolution TEM.</p>
Full article ">Figure 4
<p>Magnetization curves of AgFeS<sub>2</sub> QDs at 4 K and 298 K.</p>
Full article ">Figure 5
<p>UV–visible absorption spectra of AgFeS<sub>2</sub> QDs (red), PL excitation spectra (green) and PL emission spectra (blue) of the AgFeS<sub>2</sub> QDs dispersed in ethanol:water = 1:4.</p>
Full article ">Figure 6
<p>(<b>a</b>) The fluorescence responses of the AgFeS<sub>2</sub> QDs (2 mL) dispersed in ethanol:water = 1:4 towards different interfering substances (50 μM). Excitation wavelength: 368 nm. (<b>b</b>) A column graph of the emission changes in the presence of interfering substances (50 μM).</p>
Full article ">Figure 7
<p>(<b>a</b>) The fluorescence responses of the AgFeS<sub>2</sub> QDs (2 mL) dispersed in ethanol:water = 1:4 towards iodide from 0 to 37.5 μM (Ex = 368 nm). (<b>b</b>) The emission band at 458 nm (F<sub>0</sub>-F) versus the iodide concentration in ethanol:water = 1:4 (F and F<sub>0</sub> are the fluorescence intensity of the AgFeS<sub>2</sub> QDs in the presence and absence of I<sup>−</sup>).</p>
Full article ">Figure 8
<p>The evolution of the PL decay dynamics of the AgFeS<sub>2</sub> QDs samples in the absence (blue) and the presence (red) of 50 μΜ I<sup>−</sup> ions.</p>
Full article ">Figure 9
<p>TEM images of (<b>a</b>) the AgFeS<sub>2</sub> QDs dispersed in ethanol:water = 1:4 and (<b>b</b>) the AgFeS<sub>2</sub> QDs with the addition of 50 μΜ I<sup>−</sup> ions. (<b>c</b>) AgFeS<sub>2</sub> QDs dispersed in ethanol:water = 1:4 and maintained at room temperature for 10 h; (<b>d</b>) AgFeS<sub>2</sub> QDs dispersed in ethanol:water = 1:4 and maintained at room temperature for 24 h; (<b>e</b>) AgFeS<sub>2</sub> QDs with the addition of 50 μΜ I<sup>−</sup> ions for 10 minutes; (<b>f</b>) AgFeS<sub>2</sub> QDs with the addition of 50 μΜ I<sup>−</sup> ions for 1 h.</p>
Full article ">Figure 10
<p>The effect of ascorbic acid (50 μM) on the fluorescence emission spectra of the AgFeS<sub>2</sub> QDs dispersed in ethanol:water = 1:4.</p>
Full article ">Scheme 1
<p>Schematic illustrations of the growth of AgFeS<sub>2</sub> QDs via the one-step hot injection strategy and its application in I<sup>−</sup> sensing (ODT: octadecanethiol; OLA: oleylamine).</p>
Full article ">
16 pages, 3691 KiB  
Article
Mixed Systems of Quaternary Ammonium Foam Drainage Agent with Carbon Quantum Dots and Silica Nanoparticles for Improved Gas Field Performance
by Yongqiang Sun, Yongping Zhang, Anqi Wei, Xin Shan, Qingwang Liu, Zhenzhong Fan, Ao Sun, Lin Zhu and Lingjin Kong
Nanomaterials 2024, 14(19), 1590; https://doi.org/10.3390/nano14191590 - 1 Oct 2024
Viewed by 810
Abstract
Foam drainage agents enhance gas production by removing wellbore liquids. However, due to the ultra-high salinity environments of the Hechuan gas field (salinity up to 32.5 × 104 mg/L), no foam drainage agent is suitable for this gas field. To address this [...] Read more.
Foam drainage agents enhance gas production by removing wellbore liquids. However, due to the ultra-high salinity environments of the Hechuan gas field (salinity up to 32.5 × 104 mg/L), no foam drainage agent is suitable for this gas field. To address this challenge, we developed a novel nanocomposite foam drainage system composed of quaternary ammonium and two types of nanoparticles. This work describes the design and synthesis of a quaternary ammonium foam drainage agent and nano-engineered stabilizers. Nonylphenol polyoxyethylene ether sulfosuccinate quaternary ammonium foam drainage agent was synthesized using maleic anhydride, sodium chloroacetate, N,N-dimethylpropylenediamine, etc., as precursors. We employed the Stöber method to create hydrophobic silica nanoparticles. Carbon quantum dots were then prepared and functionalized with dodecylamine. Finally, carbon quantum dots were incorporated into the mesopores of silica nanoparticles to enhance stability. Through optimization, the best performance was achieved with a (quaternary ammonium foam drainage agents)–(carbon quantum dots/silica nanoparticles) ratio of 5:1 and a total dosage of 1.1%. Under harsh conditions (salinity 35 × 104 mg/L, condensate oil 250 cm3/m3, temperature 80 °C), the system exhibited excellent stability with an initial foam height of 160 mm, remaining at 110 mm after 5 min. Additionally, it displayed good liquid-carrying capacity (160 mL), low surface tension (27.91 mN/m), and a long half-life (659 s). These results suggest the effectiveness of nanoparticle-enhanced foam drainage systems in overcoming high-salinity challenges. Previous foam drainage agents typically exhibited a salinity resistance of no more than 25 × 104 mg/L. In contrast, this innovative system demonstrates a superior salinity tolerance of up to 35 × 104 mg/L, addressing a significant gap in available agents for high-salinity gas fields. This paves the way for future development of advanced foam systems for gas well applications with high salinity. Full article
Show Figures

Figure 1

Figure 1
<p>Reaction principle of the main foam drainage agent.</p>
Full article ">Figure 2
<p>Reaction principle of silane quaternary ammonium-salt-modified SiO<sub>2</sub>.</p>
Full article ">Figure 3
<p>Synthesis route of carbon quantum dots (CQDs).</p>
Full article ">Figure 4
<p>Preparation route of CQDs/Nano-SiO<sub>2</sub>.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) TEM images of CQDs/Nano-SiO<sub>2</sub>; (<b>c</b>,<b>d</b>) EDS scan of CQDs/Nano-SiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>Foam height and liquid holdup at different foam drainage agent concentrations.</p>
Full article ">Figure 7
<p>Half-life and surface tension at different foam drainage agent concentrations.</p>
Full article ">Figure 8
<p>Foam height and liquid holdup under different ratios.</p>
Full article ">Figure 9
<p>Half-life and surface tension under different ratios.</p>
Full article ">Figure 10
<p>Performance of foam drainage agent at different temperatures.</p>
Full article ">Figure 11
<p>Performance of foam drainage agent at different salinities.</p>
Full article ">Figure 12
<p>Performance of foam drainage agent with different amounts of condensate oil.</p>
Full article ">Figure 13
<p>Performance of foam drainage agent at different aging times.</p>
Full article ">Figure 14
<p>Effect of different foam drainage agent concentrations on surface tension.</p>
Full article ">Figure 15
<p>Comparison of half-life of foam drainage system concentrations of 1% and 1.1% in formation water.</p>
Full article ">Figure 16
<p>Comparison of half-life of foam drainage system concentrations of 1% and 1.1% in simulated water.</p>
Full article ">
11 pages, 2977 KiB  
Article
A Fluorescence Strategy Based on Guanidinylated Carbon Dots and FAM-Labeled ssDNA for Facile Detection of Lipopolysaccharide
by Zongfu Zheng, Junrong Li, Gengping Pan, Jing Wang, Yao Wang, Kai Peng, Xintian Zhang, Zhengjun Huang and Shaohuang Weng
Chemosensors 2024, 12(10), 201; https://doi.org/10.3390/chemosensors12100201 - 1 Oct 2024
Viewed by 584
Abstract
The detection of lipopolysaccharide (LPS) has important value for the monitoring of diseases such as sepsis and the impurity control of drugs. In this work, we prepared guanidinylated carbon dots (GQ-CDs) and used them to adsorb 5-carboxyfluorescein (FAM)-labeled single-stranded DNA (ssDNA) to become [...] Read more.
The detection of lipopolysaccharide (LPS) has important value for the monitoring of diseases such as sepsis and the impurity control of drugs. In this work, we prepared guanidinylated carbon dots (GQ-CDs) and used them to adsorb 5-carboxyfluorescein (FAM)-labeled single-stranded DNA (ssDNA) to become GQ-CDs/FAM-DNA, resulting in quenched FAM. The quenching efficiency of the FAM-DNA by GQ-CDs in the GQ-CDs/FAM-DNA system was 91.95%, and this quenching was stable over the long term. Upon the addition of LPS, the quenched FAM-DNA in the GQ-CDs/FAM-DNA system regained fluorescence at 520 nm. The mechanism studies found that the addition of LPS promoted the dissociation of FAM-DNA adsorbed on GQ-CDs, thereby restoring fluorescence. The degree of fluorescence recovery was closely related to the content of LPS. Under optimized conditions, the fluorescence recovery was linearly related to LPS concentrations ranging from 5 to 90 μg/mL, with a detection limit of 0.75 μg/mL. The application of this method to plasma samples and trastuzumab injections demonstrated good spiked recoveries and reproducibility. This platform, based on GQ-CDs for the adsorption and quenching of FAM-DNA, enables the detection of LPS through relatively simple mixing operations, showing excellent competitiveness for the determination of actual samples under various conditions. Full article
Show Figures

Figure 1

Figure 1
<p>The characterization of GQ-CDs. The emission spectra with different excitation wavelengths (<b>A</b>); UV–vis absorbance spectrum, with the maximum excitation and emission spectra and appearance under white and UV light (inset) (<b>B</b>); TEM, HRTEM, and size distribution (<b>C</b>); and the FTIR spectrum (<b>D</b>).</p>
Full article ">Figure 2
<p>The XPS characterization of the GQ-CDs. The wide-spectrum (<b>A</b>), C1s (<b>B</b>), N1s (<b>C</b>), and O1s (<b>D</b>) spectra.</p>
Full article ">Figure 3
<p>The interaction between GQ-CDs and FAM-DNA. The changed emission spectra of FAM-DNA with different contents of GQ-CDs (<b>A</b>), the variable QEs (<b>B</b>), the effect of time on the emission spectra of FAM-DNA (<b>C</b>), and QEs (<b>D</b>). n = 3.</p>
Full article ">Figure 4
<p>The emission spectra of FAM-DNA and GQ-CDs/FAM-DNA with and without LPS (<b>A</b>), the emission spectra of GQ-CDs with and without LPS (<b>B</b>), the zeta potential of GQ-CDs (<b>C</b>), the fluorescence anisotropy of FAM-DNA and GQ-CDs/FAM-DNA with and without LPS (<b>D</b>). n = 3.</p>
Full article ">Figure 5
<p>The detection of LPS. The emission spectra of FAM-DNA in GQ-CDs/FAM-DNA with different contents of LPS (<b>A</b>), the fitting equation of desorption rates and LPS (<b>B</b>), the specificity of this method to variable interferents (<b>C</b>). n = 3.</p>
Full article ">Scheme 1
<p>The formation of GQ-CDs/FAM-DNA and the detection of LPS through the substitution effect.</p>
Full article ">
17 pages, 5303 KiB  
Article
Fluorescent Nanocomposites of Cadmium Sulfide Quantum Dots and Polymer Matrices: Synthesis, Characterization, and Sensing Application
by Paula Méndez, Karla Ramírez, Alex Lucero, Johny Rodríguez and Betty López
Coatings 2024, 14(10), 1256; https://doi.org/10.3390/coatings14101256 - 1 Oct 2024
Viewed by 1054
Abstract
Fluorescent materials for sensing have gained attention for the visual detection of different substances as metals and pesticides for environmental monitoring. This work presents fluorescent nanocomposites in solution, film, and paper obtained without capping and stabilizing agents, coming from quantum dots of cadmium [...] Read more.
Fluorescent materials for sensing have gained attention for the visual detection of different substances as metals and pesticides for environmental monitoring. This work presents fluorescent nanocomposites in solution, film, and paper obtained without capping and stabilizing agents, coming from quantum dots of cadmium sulfide (CdS QDs) and anionic–cationic polymer matrices. Fluorescent films were formed by casting and fluorescent paper by impregnation from the solutions. The optical properties of CdS QDs in solution showed absorption between 418 and 430 nm and a maximum emission at 460 nm. TEM analysis evidenced particle size between 3 and 6 nm and diffraction patterns characteristic of CdS nanocrystals. Infrared spectra evidenced changes in the wavenumber in the fluorescent films. The band gap values (2.95–2.82 eV) suggested an application for visible transmitting film. Fluorescent solutions by UV-vis and fluorescence evidenced a chemical interaction with glyphosate standard between 1 and 100 µg/mL concentrations. The analysis of red, green, and blue color codes (RGB) evidenced a color response of the fluorescent paper at 10 and 100 µg/mL, but the fluorescent films did not show change. Nanocomposites of chitosan and pectin, in solution and on paper, exhibited a behavior “turn-on” sensor, while carboxymethylcellulose had a “turn-off” sensor. This methodology presents three fluorescent materials with potential applications in visual sensing. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) UV–vis spectra of CdS QD chitosan (Ch), carboxymethylcellulose (CMC), and pectin (Pec) nanocomposite solutions. (<b>B</b>) Photographs of the fluorescent nanocomposite solutions under visible light and a UV lamp, Blank-B, F1 (Cd:S 5:2.5 mM), F2 (Cd:S 5:5 mM), and F3 (Cd:S 10:5 mM). (<b>C</b>) UV–vis spectra of CdS QD chitosan (Ch), carboxymethylcellulose (CMC), and pectin (Pec) nanocomposite solutions treated with 100 µg/mL of Gly. (<b>D</b>) Photographs of the fluorescent nanocomposite solutions treated with 100 µg/mL of Gly (F1, F2, and F3).</p>
Full article ">Figure 2
<p>TEM images of (<b>A</b>) chitosan, (<b>B</b>–<b>D</b>) F1-CdS/chitosan nanocomposite, (<b>E</b>) CMC, (<b>F</b>–<b>H</b>) F1-CdS/CMC nanocomposite, (<b>I</b>) pectin, (<b>J</b>–<b>L</b>) F1-CdS/Pec nanocomposite (F1: 5 mM CdCl<sub>2</sub>·2.5H<sub>2</sub>O, 2.5 mM Na<sub>2</sub>S·XH<sub>2</sub>O and 1% (<span class="html-italic">w</span>/<span class="html-italic">v</span>) polymeric matrix). (<b>C</b>,<b>G</b>,<b>K</b>) were enlarged by 50%: (<b>D</b>) 1367 × 647 pixels, (<b>H</b>) 1449 × 649 pixels, and (<b>L</b>) 1757 × 755 pixels.</p>
Full article ">Figure 3
<p>FTIR spectra of (<b>A</b>) chitosan, (<b>B</b>) CMC, (<b>C</b>) and pectin nanocomposite films, and (<b>D</b>) photographs of the films under a UV lamp. Blank (polymer), F1 (Cd:S 5:2.5 mM), F2 (Cd:S 5:5 mM), and F3 (Cd:S 10:5 mM).</p>
Full article ">Figure 4
<p>UV–vis spectra of CdS QD chitosan (Ch), carboxymethylcellulose (CMC), and pectin (Pec) nanocomposite films.</p>
Full article ">Figure 5
<p>Images of the paper impregnated with the fluorescent nanocomposite solutions. On the left, the photography of the paper without Gly and the paper treated with 10 and 100 µg/mL of Gly is presented. All the images were taken under a 365 nm UV lamp. On the right, RGB codes for each paper are shown. The data were analyzed with n = 2.</p>
Full article ">
4 pages, 2506 KiB  
Correction
Correction: Geng et al. Intracellular Delivery of DNA and Protein by a Novel Cell-Permeable Peptide Derived from DOT1L. Biomolecules 2020, 10, 217
by Jingping Geng, Xiangli Guo, Lidan Wang, Richard Q. Nguyen, Fengqin Wang, Changbai Liu and Hu Wang
Biomolecules 2024, 14(10), 1199; https://doi.org/10.3390/biom14101199 - 24 Sep 2024
Viewed by 422
Abstract
In the original article [...] Full article
Show Figures

Figure 2

Figure 2
<p>Fluorescence labelled Dot1l peptide penetration in cultured cells. (<b>A</b>) Fluorescence microscopy of fluorescein isothiocyanate (FITC)-labeled Dot1l peptide with different concentrations. (<b>B</b>) Quantization analysis of FITC-labeled Dot1l peptide corresponding to fluorescence microscopy with different concentration measurements. The statistical analysis is shown in Supplementary Figure S1A. (<b>C</b>) Fluorescence microscopy of FITC-labeled Dot1l peptide with or without trypan blue incubation in the DMSO-pretreated or control group. (<b>D</b>) FITC-labeled Dot1l peptide penetration in different cell lines (MCF7, B16, HSC-T6, Caski, and HepG2) with or without DMSO pretreatment. The statistical analysis is shown in Supplementary Figure S1B,C. Cell lysate fluorescence intensity was adjusted by protein concentration examined by Bradford assay.</p>
Full article ">Figure 3
<p>Effects of Dot1l peptide penetration under different conditions. (<b>A</b>) Fluorescence microscopy of FITC-labeled Dot1l peptide at 37 °C and 4 °C. (<b>B</b>) Quantization of Dot1l peptide penetration in MCF7 cells at different temperatures. Data are presented as means ± SEM (<span class="html-italic">n</span> = 3). The statistical analysis is shown in Supplementary Figure S1D. (<b>C</b>) Fluorescence microscopy of different inhibitors’ exposure on Dot1l peptide penetration. (<b>D</b>) Suppression of different inhibitors’ exposure on Dot1l peptide penetration. Data are presented as means ± SEM (<span class="html-italic">n</span> = 3). The statistical analysis is shown in Supplementary Figure S1E. Cell lysate fluorescence intensity was adjusted by protein concentration examined by Bradford assay.</p>
Full article ">
12 pages, 843 KiB  
Article
Advances of the Holographic Technique to Test the Basic Properties of the Thin-Film Organics: Refractivity Change and Novel Mechanism of the Nonlinear Attenuation Prediction
by Natalia Kamanina
Polymers 2024, 16(18), 2645; https://doi.org/10.3390/polym16182645 - 19 Sep 2024
Viewed by 586
Abstract
A large number of the thin-film organic structures (polyimides, 2-cyclooctylarnino-5-nitropyridine, N-(4-nitrophenyl)-(L)-prolinol, 2-(n-Prolinol)-5-nitropyridine) sensitized with the different types of the nano-objects (fullerenes, carbon nanotubes, quantum dots, shungites, reduced graphene oxides) are presented, which are studied using the holographic technique under the Raman–Nath diffraction conditions. [...] Read more.
A large number of the thin-film organic structures (polyimides, 2-cyclooctylarnino-5-nitropyridine, N-(4-nitrophenyl)-(L)-prolinol, 2-(n-Prolinol)-5-nitropyridine) sensitized with the different types of the nano-objects (fullerenes, carbon nanotubes, quantum dots, shungites, reduced graphene oxides) are presented, which are studied using the holographic technique under the Raman–Nath diffraction conditions. Pulsed laser irradiation testing of these materials predicts a dramatic increase of the laser-induced refractive index, which is in several orders of the magnitude greater compared to pure materials. The estimated nonlinear refraction coefficients and the cubic nonlinearities for the materials studied are close to or larger than those known for volumetric inorganic crystals. The role of the intermolecular charge transfer complex formation is considered as the essential in the refractivity increase in nano-objects-doped organics. As a new idea, the shift of charge from the intramolecular donor fragment to the intermolecular acceptors can be proposed as the development of Janus particles. The energy losses via diffraction are considered as an additional mechanism to explain the nonlinear attenuation of the laser beam. Full article
(This article belongs to the Special Issue Advanced Polymer Nanocomposites III)
Show Figures

Figure 1

Figure 1
<p>The proposed extended scheme of intermolecular charge transfer using organic conjugated materials, where the introduced intermolecular doping nano-object has an electron affinity significantly greater than the intramolecular acceptor of the matrix system.</p>
Full article ">
12 pages, 8166 KiB  
Article
Paper-Based Fluorescent Sensor for Rapid Multi-Channel Detection of Tetracycline Based on Graphene Quantum Dots Coated with Molecularly Imprinted Polymer
by Linzhe Wang, Jingfang Hu, Wensong Wei, Yu Song, Yansheng Li, Guowei Gao, Chunhui Zhang and Fangting Fu
Polymers 2024, 16(17), 2540; https://doi.org/10.3390/polym16172540 - 8 Sep 2024
Viewed by 3442
Abstract
In this paper, we developed a paper-based fluorescent sensor using functional composite materials composed of graphene quantum dots (GQDs) coated with molecularly imprinted polymers (MIPs) for the selective detection of tetracycline (TC) in water. GQDs, as eco-friendly fluorophores, were chemically grafted onto the [...] Read more.
In this paper, we developed a paper-based fluorescent sensor using functional composite materials composed of graphene quantum dots (GQDs) coated with molecularly imprinted polymers (MIPs) for the selective detection of tetracycline (TC) in water. GQDs, as eco-friendly fluorophores, were chemically grafted onto the surface of paper fibers. MIPs, serving as the recognition element, were then wrapped around the GQDs via precipitation polymerization using 3-aminopropyltriethoxysilane (APTES) as the functional monomer. Optimal parameters such as quantum dot concentration, grafting time, and elution time were examined to assess the sensor’s detection performance. The results revealed that the sensor exhibited a linear response to TC concentrations in the range of 1 to 40 µmol/L, with a limit of detection (LOD) of 0.87 µmol/L. When applied to spiked detection in actual water samples, recoveries ranged from 103.3% to 109.4%. Overall, this paper-based fluorescent sensor (MIPs@GQDs@PAD) shows great potential for portable, multi-channel, and rapid detection of TC in water samples in the future. Full article
(This article belongs to the Special Issue Functional Graphene–Polymer Composites)
Show Figures

Figure 1

Figure 1
<p>A schematic illustration of the fabrication process of the MIPs@GQDs@PAD.</p>
Full article ">Figure 2
<p>A schematic illustration of the detection process of the MIPs@GQDs@PAD for TC.</p>
Full article ">Figure 3
<p>Fluorescence spectra of (<b>A</b>) MIPs@GQD emission spectra in the presence and absence of TC; (<b>B</b>) excitation and emission spectra of MIPs@GQDs@PAD.</p>
Full article ">Figure 4
<p>Fluorescence characterization of (<b>A</b>) MIPs@GQDs@PAD after elution and (<b>B</b>) MIPs@GQDs@PAD after readsorption of TC.</p>
Full article ">Figure 5
<p>SEM characterization of (<b>A</b>) bare paper, (<b>B</b>) GQDs@PAD, and (<b>C</b>) MIPs@GQDs@PAD.</p>
Full article ">Figure 6
<p>FT-IR spectra of bare paper, GQDs@PAD, and MIPs@GQDs@PAD before and after elution.</p>
Full article ">Figure 7
<p>Effect of (<b>A</b>) GQD grafting time and (<b>B</b>) amount of GQDs.</p>
Full article ">Figure 8
<p>Effect of (<b>A</b>) pH and (<b>B</b>) amount of functional monomer.</p>
Full article ">Figure 9
<p>Effect of (<b>A</b>) different eluents and (<b>B</b>) number of elution.</p>
Full article ">Figure 10
<p>The (<b>A</b>) fluorescence emission spectra of MIPs@GQDs@PAD under different TC concentrations and the (<b>B</b>) calibration curve and linear equation.</p>
Full article ">Figure 11
<p>The (<b>A</b>) reproducibility and (<b>B</b>) selectivity of MIPs@GQDs@PAD.</p>
Full article ">
17 pages, 4875 KiB  
Article
Carbon Dots-Mediated Photodynamic Treatment Reduces Postharvest Senescence and Decay of Grapes by Regulating the Antioxidant System
by Zhi-Jing Ni, Ying Xue, Wei Wang, Juan Du, Kiran Thakur, Wen-Ping Ma and Zhao-Jun Wei
Foods 2024, 13(17), 2717; https://doi.org/10.3390/foods13172717 - 27 Aug 2024
Viewed by 720
Abstract
Grapes are susceptible to mold and decay during postharvest storage, and developing new technologies to extend their storage period has important application value. Photodynamic technology (PDT) in concurrence with carbon dots (CDs) proposes an innovative and eco-friendly preservation strategy. We examined the effects [...] Read more.
Grapes are susceptible to mold and decay during postharvest storage, and developing new technologies to extend their storage period has important application value. Photodynamic technology (PDT) in concurrence with carbon dots (CDs) proposes an innovative and eco-friendly preservation strategy. We examined the effects of carbon dots combined with photodynamic treatment on postharvest senescence and antioxidant system of table grape. The compounding of photodynamic technology with a 0.06 g L−1 CDs solution could possibly extend the postharvest storage period of grape berries. Through this strategy, we achieved a decreased rate of fruit rotting and weight loss alongside the delayed deterioration of fruit firmness, soluble solids, and titratable acid. As paired with photodynamic technology, CDs considerably decreased the postharvest storage loss of phenols, flavonoids, and reducing sugars as compared to the control group. Concurrently, it remarkably postponed the build-up of hydrogen peroxide (H2O2), superoxide anion (O2∙−), and malondialdehyde (MDA); elevated the levels of reduced ascorbic acid (AsA) and reduced glutathione (GSH); lowered the levels of dehydroascorbic acid (DHA) and oxidized glutathione (GSSG); raised the ratios of AsA/DHA and GSSH/GSSG; encouraged the activities of superoxide dismutase (SOD) and phenylalanine ammonia-lyase (PAL); and inhibited the activities of polyphenol oxidase (PPO) and lipoxygenase (LOX). Furthermore, it enhanced the iron reduction antioxidant capacity (FRAP) and DPPH radical scavenging capacity of grape berries. CDs combined with photodynamic treatment could efficiently lessen postharvest senescence and decay of grape berry while extending the storage time. Full article
(This article belongs to the Special Issue Postharvest Storage and Preservation of Fruits and Vegetables)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of CDs-mediated PDT-treated grapes. (<b>A</b>) Composite diagram of CDs and (<b>B</b>) flow diagram of grape treatment under light.</p>
Full article ">Figure 2
<p>Effects of different treatment conditions on the overall grape appearance (<b>A</b>) and decay rate (<b>B</b>). P-L-: control group; P+L+: spray 0.06 g L<sup>−1</sup> CDs solution and 450 nm LED light for 5 min; P+L-: CDs treatment separately (0.06 g L<sup>−1</sup>) without illumination; P-L+: 450 nm LED alone for 5 min, no CDs treatment. After different treatments, the Kyoho grapes were stored at 25 °C for 12 days. The data are expressed as the mean ± SD (standard deviation) (n = 3). Different letters indicate significant difference between treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 3
<p>Effect of CDs-mediated photodynamic on the physical properties of postharvest grapes. The rotten fruit rate of fresh grape during storage period; weight loss rate; firmness; and TSS content, TA content, and TSS/TA (<b>A</b>–<b>F</b>). Treatment group: 0.06 g L<sup>−1</sup> CDs solution and blue light irradiation for 5 min; control group: distilled water instead of treatment and storage at 25 °C for 12 days. * and ** represent significant differences between the sample point treatment and control groups, where * represents <span class="html-italic">p</span> &lt; 0.05, and ** represents <span class="html-italic">p</span> &lt; 0.01, and data are presented as mean ± SD (standard deviation) (n = 3).</p>
Full article ">Figure 4
<p>Effect of CDs-mediated PDT of O<sub>2</sub><sup>∙−</sup>content (<b>A</b>), H<sub>2</sub>O<sub>2</sub> content (<b>B</b>), and MDA content (<b>C</b>) in grape skin and pulp. Treatment group: 0.06 g L<sup>−1</sup> CDs solution and blue light irradiation for 5 min; control group: distilled water instead of treatment and storage at 25 °C for 12 days. * and ** represent significant differences between the sample point treatment and control groups, where * represents <span class="html-italic">p</span> &lt; 0.05, and ** represents <span class="html-italic">p</span> &lt; 0.01, and data are presented as mean ± SD (standard deviation) (n = 3).</p>
Full article ">Figure 5
<p>Effect of CDs-mediated PDT of total phenols content, flavonoid, reducing sugar, AsA content, GSH content, DHA content, GSSG content, AsA/DHA, and GSH/GSSG in grape (<b>A</b>–<b>I</b>). Treatment group: 0.06 g L<sup>−1</sup> CDs solution and blue light irradiation for 5 min; control group: distilled water instead of treatment and storage at 25 °C for 12 days. * and ** represent significant differences between the sample point treatment and control groups, where * represents <span class="html-italic">p</span> &lt; 0.05, and ** represents <span class="html-italic">p</span> &lt; 0.01, and data are presented as mean ± SD (standard deviation) (n = 3).</p>
Full article ">Figure 6
<p>Effect of CDs-mediated PDT of SOD activity, PAL activity, PPO activity, LOX activity, DPPH, and FRAP in grape (<b>A</b>–<b>F</b>); carbon dots mediated the relationship between quality indexes, antioxidant indexes, and total antioxidant capacity indexes in grape skin (<b>G</b>) and pulp (<b>H</b>) in photodynamic treatment. The edge width corresponds to Mantel’s r-value, the color represents statistical significance, and the pairwise correlation of these variables is represented by the color gradient of the Pearson correlation coefficient. Treatment group: 0.06 g L <sup>− 1</sup> CDs solution and blue light irradiation for 5 min; control group: distilled water instead of treatment and storage at 25 °C for 12 days. * and ** represent significant differences between the sample point treatment and control groups, where * represents <span class="html-italic">p</span> &lt; 0.05, and ** represents <span class="html-italic">p</span> &lt; 0.01, and data are presented as mean ± SD (standard deviation) (n = 3).</p>
Full article ">
Back to TopTop