[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
remove_circle_outline

Search Results (2,105)

Search Parameters:
Keywords = organic dye

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
26 pages, 2814 KiB  
Review
Recent Advances of Strategies and Applications in Aptamer-Combined Metal Nanocluster Biosensing Systems
by Ki-Beom Kim, Sang-Ho Kim and Seung-Min Yoo
Biosensors 2024, 14(12), 625; https://doi.org/10.3390/bios14120625 (registering DOI) - 18 Dec 2024
Abstract
Metal nanoclusters (NCs) are promising alternatives to organic dyes and quantum dots. These NCs exhibit unique physical and chemical properties, such as fluorescence, chirality, magnetism and catalysis, which contribute to significant advancements in biosensing, biomedical diagnostics and therapy. Through adjustments in composition, size, [...] Read more.
Metal nanoclusters (NCs) are promising alternatives to organic dyes and quantum dots. These NCs exhibit unique physical and chemical properties, such as fluorescence, chirality, magnetism and catalysis, which contribute to significant advancements in biosensing, biomedical diagnostics and therapy. Through adjustments in composition, size, chemical environments and surface ligands, it is possible to create NCs with tunable optoelectronic and catalytic activity. This review focuses on the integration of aptamers with metal NCs, detailing molecular detection strategies that utilise the effect of aptamers on optical signal emission of metal NC-based biosensing systems. This review also highlights recent advancements in biosensing and biomedical applications, as well as illustrative case studies. To conclude, the strengths, limitations, current challenges and prospects for metal NC-based systems were examined. Full article
(This article belongs to the Special Issue Biomaterials for Biosensing Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration highlighting the features of aptamer and metal nanoclusters.</p>
Full article ">Figure 2
<p>Schematic illustration of three strategies for signal changes induced by combining aptamer with metal NCs. MOF, metal-organic framework; COF, covalent-organic framework.</p>
Full article ">Figure 3
<p>Sensing strategy based on signal emission changes in aptamer-linked metal NCs. (<b>A</b>) Detection of urea, ATP, estradiol using NC-loaded COF and aptamer. This system is dual-mode SERS and RRS sensor. Reproduced with permission from [<a href="#B93-biosensors-14-00625" class="html-bibr">93</a>]. Copyright 2020, Elsevier. (<b>B</b>) Detection of T-2 toxin using PAA@Arg@ATT-AuNCs NPs and aptamer–PDDA complex. This system used FRET between PAA@Arg@ATT-AuNCs (fluorescence donor) and AuNPs (energy receptor). Reproduced with permission from [<a href="#B86-biosensors-14-00625" class="html-bibr">86</a>]. Copyright 2020, Elsevier. (<b>C</b>) Detection of <span class="html-italic">Salmonella typhimurium</span> using AuNCs@aptamer and TMB. This system enables simultaneous binding of bacteria to both the aptamer@AuNCs and TMB, facilitating peroxidase-like activity due to the increased proximity of these interactions. Reproduced with permission from [<a href="#B91-biosensors-14-00625" class="html-bibr">91</a>]. Copyright 2020, Elsevier. (<b>D</b>) Detection of two different mycotoxins (aflatoxin B1 and zearalenone) using FRET between the AuNCs and WS<sub>2</sub> quencher. Reproduced with permission from [<a href="#B63-biosensors-14-00625" class="html-bibr">63</a>]. Copyright 2019, American Chemical Society. NC, nanocluster; ATP, adenosine triphosphate; COF, covalent-organic framework; PAA, polyacrylic acid; ATT, 6-aza-2-thiothymine; PDDA, poly (diallyldimethylammonium chloride); TMB, tetramethylbenzidine.</p>
Full article ">Figure 4
<p>Sensing strategy based on signal changes in metal NC produced by aptamer-linked DNA template. (<b>A</b>) Detection of Pb<sup>2+</sup> using a scaffold of the AgNC formation template fused with aptamer to form G-quadruplex structure in the presence of target. Reproduced with permission from [<a href="#B99-biosensors-14-00625" class="html-bibr">99</a>]. Copyright 2018, Elsevier. (<b>B</b>) Detection of kanamycin using the scaffolds consisting of two split aptamer and Cu/Ag bimetal NC formation templates. Reproduced with permission from [<a href="#B103-biosensors-14-00625" class="html-bibr">103</a>]. Copyright 2022, Elsevier. (<b>C</b>) Detection of T-2 toxin using a scaffold containing an aptamer, a T-linker and an AgNC template. This system also used FRET between MoS<sub>2</sub> nanosheets (fluorescence acceptor) and the aptamer–AgNCs (fluorescence donor). Reproduced with permission from [<a href="#B64-biosensors-14-00625" class="html-bibr">64</a>]. Copyright 2018, Elsevier. (<b>D</b>) Detection of ZEN using a scaffold consisting of an AgNC template, an aptamer and a G-rich domain. This system uses of FRET between the aptamer–AgNCs and porous Fe<sub>3</sub>O<sub>4</sub>/C acting on quenching of fluorescence and the easy separation. Reproduced with permission from [<a href="#B100-biosensors-14-00625" class="html-bibr">100</a>]. Copyright 2021, Elsevier. (<b>E</b>) Detection of three different tumour biomarkers (mucin 1, carcinoembryonic antigen and cancer antigen 125), using a scaffold consisting of the same NC nucleation sequence and different aptamer sequences exhibiting different emission wavelengths for the detection of three molecules. This system used FRET between Ag/Au bimetallic NCs (donor) and GOx nanosheets (quencher). Reproduced with permission from [<a href="#B59-biosensors-14-00625" class="html-bibr">59</a>]. Copyright 2018, Elsevier. (<b>F</b>) Detection of MUC1 using a scaffold consisting of C-rich template and aptamer with G-rich sequence at the end. Reproduced with permission from [<a href="#B112-biosensors-14-00625" class="html-bibr">112</a>]. Copyright 2019, Elsevier. NC, nanocluster; CA125, cancer antigen 125; CEA, carcinoembryonic antigen; MUC1, mucin 1; APT, aptamer.</p>
Full article ">Figure 5
<p>Sensing strategy based on signal changes in metal NCs induced by aptamer–DNA template hybridisation. (<b>A</b>) Detection of two different bacterial cells (<span class="html-italic">Staphylococcus aureus</span> and <span class="html-italic">Escherichia coli</span>) using AgNC bound with hybrid DNA of NC scaffold and bacteria-specific aptamer. This system used the antibacterial effect of AgNC and enhanced AgNC fluorescence via electrospinning to PLA, forming nanofilms. Reproduced with permission from [<a href="#B107-biosensors-14-00625" class="html-bibr">107</a>]. Copyright 2021, American Chemical Society. (<b>B</b>) Detection of ZEN using dual-signal amplification mechanism based on TdT amplification and CuNC fluorescence enhancement. Reproduced with permission from [<a href="#B76-biosensors-14-00625" class="html-bibr">76</a>]. Copyright 2024, Elsevier. (<b>C</b>) Detection of ochratoxin A using aptamer serving as both the recognition and quenching reagent. This system used scaffold sequences screened for emitting or quenching fluorescence. Reproduced with permission from [<a href="#B106-biosensors-14-00625" class="html-bibr">106</a>]. Copyright 2023, Elsevier. NC, nanocluster; PLA, polylactic acid; ZEN, zearalenone; SMB, streptavidin-coated magnetic bead; TdT, terminal deoxynucleotidyl transferase; OTA, ochratoxin A.</p>
Full article ">
13 pages, 4035 KiB  
Communication
Use of Laccase Enzymes as Bio-Receptors for the Organic Dye Methylene Blue in a Surface Plasmon Resonance Biosensor
by Araceli Sánchez-Álvarez, Gabriela Elizabeth Quintanilla-Villanueva, Osvaldo Rodríguez-Quiroz, Melissa Marlene Rodríguez-Delgado, Juan Francisco Villarreal-Chiu, Analía Sicardi-Segade and Donato Luna-Moreno
Sensors 2024, 24(24), 8008; https://doi.org/10.3390/s24248008 - 15 Dec 2024
Viewed by 458
Abstract
Methylene blue is a cationic organic dye commonly found in wastewater, groundwater, and surface water due to industrial discharge into the environment. This emerging pollutant is notably persistent and can pose risks to both human health and the environment. In this study, we [...] Read more.
Methylene blue is a cationic organic dye commonly found in wastewater, groundwater, and surface water due to industrial discharge into the environment. This emerging pollutant is notably persistent and can pose risks to both human health and the environment. In this study, we developed a Surface Plasmon Resonance Biosensor employing a BK7 prism coated with 3 nm chromium and 50 nm of gold in the Kretschmann configuration, specifically for the detection of methylene blue. For the first time, laccases immobilized on a gold surface were utilized as bio-receptors for this organic dye. The enzyme was immobilized using carbodiimide bonds with EDC/NHS crosslinkers, allowing for the analysis of samples with minimal preparation. The method demonstrated validation with a limit of detection (LOD) of 4.61 mg L−1 and a limit of quantification (LOQ) of 15.37 mg L−1, a working range of 0–100 mg L−1, and an R2 value of 0.9614 during real-time analysis. A rainwater sample spiked with methylene blue yielded a recovery rate of 122.46 ± 4.41%. The biosensor maintained a stable signal over 17 cycles and remained effective for 30 days at room temperature. Full article
(This article belongs to the Section Biosensors)
Show Figures

Figure 1

Figure 1
<p>Chemical structure of methylene blue.</p>
Full article ">Figure 2
<p>Possible recognition process and first step of degradation of methylene blue by laccases.</p>
Full article ">Figure 3
<p>Immobilization process of laccases on the thin chromium–gold film chip. In the step 1, alkanethiols are added to the thin gold surface. In step 2, the EDC is added, forming an unstable intermediate. In step 3, the NHS is added, creating a sulfo-NHS ester. In step 4, NHS is replaced by the laccase through an amide bond.</p>
Full article ">Figure 4
<p>Assembly of the prism and the chip on the SPR equipment: (<b>a</b>) assembly of the prism, the chip with a thin gold film with the immobilized laccases, the prism and other components. (<b>b</b>) Set up of the prism, sample cell, chip with immobilized laccases and the other on the SPR equipment.</p>
Full article ">Figure 5
<p>Reflectance spectra obtained by angular sweep.</p>
Full article ">Figure 6
<p>Immobilization process of laccass from <span class="html-italic">Rhus vernicifera</span> in real-time by SPR.</p>
Full article ">Figure 7
<p>FTIR analysis of different stages of laccase immobilization.</p>
Full article ">Figure 8
<p>(<b>a</b>) SPR analysis of stocks with different concentrations of methylene blue. (<b>b</b>) Calibration curve and equation of a straight line.</p>
Full article ">Figure 9
<p>Comparison of the intensity of reflectance of solutions of methylene blue at day 1 and day 30.</p>
Full article ">
28 pages, 12272 KiB  
Article
New Derivatives of Chalcones, Chromenes, and Stilbenoids, Complexed with Methyl-β-Cyclodextrin with Antioxidant Properties and Antibacterial Synergism with Antibiotics
by Igor D. Zlotnikov, Sergey S. Krylov, Natalya G. Belogurova, Alexander N. Blinnikov, Victor E. Kalugin and Elena V. Kudryashova
Biophysica 2024, 4(4), 667-694; https://doi.org/10.3390/biophysica4040044 - 13 Dec 2024
Viewed by 285
Abstract
Cyclodextrins (CDs) are natural cyclic oligosaccharides with the ability to form inclusion complexes with various organic substances. In this paper, we investigate the potential of CD complex formation to enhance the antibacterial activity and antioxidant properties of poorly soluble bioactive agents, such as [...] Read more.
Cyclodextrins (CDs) are natural cyclic oligosaccharides with the ability to form inclusion complexes with various organic substances. In this paper, we investigate the potential of CD complex formation to enhance the antibacterial activity and antioxidant properties of poorly soluble bioactive agents, such as chalcones, chromenes, stilbenoids and xanthylium derivatives, serving as potential adjuvants, in comparison with standard antiseptics. The interaction of these bioactive agents with the hydrophobic pocket of methyl-β-cyclodextrin (MCD) was confirmed using spectroscopic methods such as UV-vis, FTIR, 1H and 13C NMR, mass-spectrometry. CD-based delivery system allows for combining multiple active agents, improving solubility, antibacterial efficacy by enhancing penetration into target bacterial cells (E. coli selectivity demonstrated via confocal microscopy). Novel compounds of chalcones and stilbenoids derivatives additionally enhance efficacy by inhibiting bacterial efflux pumps, increasing membrane permeability, and inhibiting bacterial enzymes, and showed a synergy when used in combination with metronidazole. The intricate relationship between the structural characteristics and functional properties of chalcones and stilbenoids in terms of their antibacterial and antioxidative capabilities is revealed. The substituents within aromatic rings significantly influence this activity, where position of electron-donating methoxy groups playing a crucial role. Among chalcones, stilbenoids, ana xanthyliums, the compounds caring a benzodioxol ring, analogous to natural bioactive compounds like apiol, dillapiol, and myristicin, emerge as prominent antibacterial activity. To explore the possibility to create theranostic formulations, we used fluorescent markers to visualize target cells, antiseptics to provide antibacterial activity, and bioactive agents as chalcones acting as adjuvants. Additionally, new antioxidant compounds were found such as Xanthylium derivative (R351) and chromene derivative: 1-methyl-3-(2-amino-3-cyano-7-methoxychromene-4-yl)-pyridinium methanesulfate: the pronounced antioxidant properties of these substances are observed comparable to quercetin in the efficiency. Rhodamine 6G, gentian violet, and Congo Red exhibit good antioxidant properties, although their activity is an order of magnitude lower than that of quercetin. However, they have remarkable potential due to their multifaceted nature, including the ability to visualize target cells. The most effective theranostic formulation is the combination of the antibiotic (metronidazole) + dye/fluorophore (methylene blue/rhodamine 6G) for visualization of target cells + adjuvant (chalcones or xanthylium derivatives) for antiinflammation effect. This synergistic combination, results in a promising theranostic formulation for treating bacterial infections, with enhanced efficiency, selectivity and minimizing side effects. Full article
Show Figures

Figure 1

Figure 1
<p>UV/vis absorption spectra of the investigated dyes and chalcones in free form in a buffer solution and complexed with methyl-cyclodextrin (MCD): (<b>a</b>) Congo Red 8 µM; (<b>b</b>) Methylene blue 15 µM; (<b>c</b>) Malachite green 10 µM; (<b>d</b>) R351-ClO<sub>4</sub><sup>−</sup> 2.5 µM; (<b>e</b>) sample 9–50 µM; (<b>f</b>) Brilliant green 35 µM, toluidine blue 5 µM, gentian violet 3 µM, sudan III 7 µM. PBS (0.01M, pH 7.4). DMSO could be added to enhance the solubility of free samples. T = 37 °C.</p>
Full article ">Figure 1 Cont.
<p>UV/vis absorption spectra of the investigated dyes and chalcones in free form in a buffer solution and complexed with methyl-cyclodextrin (MCD): (<b>a</b>) Congo Red 8 µM; (<b>b</b>) Methylene blue 15 µM; (<b>c</b>) Malachite green 10 µM; (<b>d</b>) R351-ClO<sub>4</sub><sup>−</sup> 2.5 µM; (<b>e</b>) sample 9–50 µM; (<b>f</b>) Brilliant green 35 µM, toluidine blue 5 µM, gentian violet 3 µM, sudan III 7 µM. PBS (0.01M, pH 7.4). DMSO could be added to enhance the solubility of free samples. T = 37 °C.</p>
Full article ">Figure 2
<p>FTIR spectra of the investigated «drug candidates» in free form in a buffer solution and complexed with methyl-cyclodextrin (MCD): (<b>a</b>) chalcone (9), (<b>b</b>) 1-methyl-3-(2-amino-3-cyano-7-methoxychromene-4-yl)-pyridinium methanesulfate (17). PBS (0.01M, pH 7.4). T = 37 °C.</p>
Full article ">Figure 3
<p><sup>1</sup>H NMR spectra of the «drug candidate» sample 9: (<b>a</b>) in free form in d<sub>6</sub>-DMSO; (<b>b</b>) complexed with methyl-cyclodextrin (MCD) (1:5 mol/mol) in D<sub>2</sub>O with predicted peak correlations. T = 25 °C. (<b>c</b>) The proposed structure of the chalcone 9—MCD complex obtained during computer modeling. Carbon atoms are indicated in green (MCD) and blue (the guest molecule, compound <b>9</b>); oxygen atoms are indicated in red. hydrogen—white, sulfur—yellow, nitrogen—blue. The purple sphere is Na<sup>+</sup>. The simulation was performed using the PyMOL program. (<b>d</b>) Schematic cyclodextrin torus representation.</p>
Full article ">Figure 3 Cont.
<p><sup>1</sup>H NMR spectra of the «drug candidate» sample 9: (<b>a</b>) in free form in d<sub>6</sub>-DMSO; (<b>b</b>) complexed with methyl-cyclodextrin (MCD) (1:5 mol/mol) in D<sub>2</sub>O with predicted peak correlations. T = 25 °C. (<b>c</b>) The proposed structure of the chalcone 9—MCD complex obtained during computer modeling. Carbon atoms are indicated in green (MCD) and blue (the guest molecule, compound <b>9</b>); oxygen atoms are indicated in red. hydrogen—white, sulfur—yellow, nitrogen—blue. The purple sphere is Na<sup>+</sup>. The simulation was performed using the PyMOL program. (<b>d</b>) Schematic cyclodextrin torus representation.</p>
Full article ">Figure 4
<p>ABTS antioxidant test of the “drug candidates” complexed with methyl-cyclodextrin (MCD) (1:5 mol/mol): (<b>a</b>) Dyes and fluorophores; (<b>b</b>) Chalcone and stilbene derivatives. PBS (0.01M, pH 7.4). T = 37 °C.</p>
Full article ">Figure 4 Cont.
<p>ABTS antioxidant test of the “drug candidates” complexed with methyl-cyclodextrin (MCD) (1:5 mol/mol): (<b>a</b>) Dyes and fluorophores; (<b>b</b>) Chalcone and stilbene derivatives. PBS (0.01M, pH 7.4). T = 37 °C.</p>
Full article ">Figure 5
<p>The structures of the drug candidates in the relationships with antioxidant and antibacterial properties. The circles highlight significant fragments of molecules.</p>
Full article ">Figure 6
<p>Confocal laser scanning microscopy images of (<b>a</b>,<b>b</b>) <span class="html-italic">E. coli</span> cells and (<b>c</b>,<b>d</b>) <span class="html-italic">Lactobacilli</span> cells, stained with R6G (10 µg/mL) in free form or complexed with MCD (100 µg/mL). λexci, max = 488 nm, λemi = 530–580 nm. The scale bar is 20 µm.</p>
Full article ">Scheme 1
<p>Synthesis scheme of compound <b>17</b>.</p>
Full article ">Scheme 2
<p>Synthesis scheme of Xanthylium derivatives R351 salts compounds.</p>
Full article ">Scheme 3
<p>Synthesis scheme of the chalcone and stilbenoid derivatives.</p>
Full article ">
10 pages, 1710 KiB  
Article
Quantum Chemical Determination of Molecular Dye Candidates for Non-Invasive Bioimaging
by Remy R. Cron, Jordan South and Ryan C. Fortenberry
Molecules 2024, 29(24), 5860; https://doi.org/10.3390/molecules29245860 - 12 Dec 2024
Viewed by 381
Abstract
Molecular dyes containing carbazole-based π bridges and/or julolidine-based donors should be promising molecules for intense SWIR emission with potential application to molecular bioimaging. This study stochastically analyzes the combinations of more than 250 organic dyes constructed within the D-π-D (or equivalently [...] Read more.
Molecular dyes containing carbazole-based π bridges and/or julolidine-based donors should be promising molecules for intense SWIR emission with potential application to molecular bioimaging. This study stochastically analyzes the combinations of more than 250 organic dyes constructed within the D-π-D (or equivalently D-B-D) motif. These dyes are built from 22 donors (D) and 14 π bridges (B) and are computationally examined using density functional theory (DFT). The DFT computations provide optimized geometries from which the excited state transition wavelengths and associated oscillator strengths and orbital overlaps are computed. While absorption is used as a stand-in for emission, the longer the absorption wavelength, the longer the emission should be as well for molecules of this type. Nearly 100 novel dyes reported in this work have electronic absorptions at or beyond 1200 nm, opening the possibility for future synthesis and experimental characterization of new molecular dyes with promising properties for bioimaging. Full article
(This article belongs to the Section Physical Chemistry)
Show Figures

Figure 1

Figure 1
<p>The wavelengths computed for the constructed D-<math display="inline"><semantics> <mi>π</mi> </semantics></math>-D (D-B-D) dyes. The x-axis is largely arbitrary as the dyes are sorted by wavelength.</p>
Full article ">Figure 2
<p>The orbitals involved in the 7D-3B-7D HOMO-LUMO transition.</p>
Full article ">Figure 3
<p>Chemical structures for the donors utilized in this work. Note the red noble gas (argon) atoms used as placeholders for where the linkages are made.</p>
Full article ">Figure 4
<p>Chemical structures for the <math display="inline"><semantics> <mi>π</mi> </semantics></math> bridges utilized in this work with red noble gas atom placeholders. Since there are two connection sites in these molecules, different noble gas atom placeholder (argon &amp; krypton) are used.</p>
Full article ">
17 pages, 6521 KiB  
Article
Rational Fabrication of Ag2S/g-C3N4 Heterojunction for Photocatalytic Degradation of Rhodamine B Dye Under Natural Solar Radiation
by Ali Alsalme, Ahmed Najm, Nagy N. Mohammed, M. F. Abdel Messih, Ayman Sultan and Mohamed Abdelhay Ahmed
Catalysts 2024, 14(12), 914; https://doi.org/10.3390/catal14120914 - 11 Dec 2024
Viewed by 530
Abstract
Near-infrared light-triggered photocatalytic water treatment has attracted significant attention in recent years. In this novel research, rational sonochemical fabrication of Ag2S/g-C3N4 nanocomposites with various compositions of Ag2S (0–25) wt% was carried out to eliminate hazardous rhodamine [...] Read more.
Near-infrared light-triggered photocatalytic water treatment has attracted significant attention in recent years. In this novel research, rational sonochemical fabrication of Ag2S/g-C3N4 nanocomposites with various compositions of Ag2S (0–25) wt% was carried out to eliminate hazardous rhodamine B dye in a cationic organic pollutant model. g-C3N4 sheets were synthesized via controlled thermal annealing of microcrystalline urea. However, black Ag2S nanoparticles were synthesized through a precipitation-assisted sonochemical route. The chemical interactions between various compositions of Ag2S and g-C3N4 were carried out in an ultrasonic bath with a power of 300 W. XRD, PL, DRS, SEM, HRTEM, mapping, BET, and SAED analysis were used to estimate the crystalline, optical, nanostructure, and textural properties of the solid specimens. The coexistence of the diffraction peaks of g-C3N4 and Ag2S implied the successful production of Ag2S/g-C3N4 heterojunctions. The band gap energy of g-C3N4 was exceptionally reduced from 2.81 to 1.5 eV with the introduction of 25 wt% of Ag2S nanoparticles, implying the strong absorbability of the nanocomposites to natural solar radiation. The PL signal intensity of Ag2S/g-C3N4 was reduced by 40% compared with pristine g-C3N4, implying that Ag2S enhanced the electron–hole transportation and separation. The rate of the photocatalytic degradation of rhodamine B molecules was gradually increased with the introduction of Ag2S on the g-C3N4 surface and reached a maximum for nanocomposites containing 25 wt% Ag2S. The radical trapping experiments demonstrated the principal importance of reactive oxygen species and hot holes in destroying rhodamine B under natural solar radiation. The charge transportation between Ag2S and g-C3N4 semiconductors proceeded through the type I straddling scheme. The enriched photocatalytic activity of Ag2S/g-C3N4 nanocomposites resulted from an exceptional reduction in band gap energy and controlling the electron–hole separation rate with the introduction of Ag2S as an efficient photothermal photocatalyst. The novel as-synthesized nanocomposites are considered a promising photocatalyst for destroying various types of organic pollutants under low-cost sunlight radiation. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>XRD of g-C<sub>3</sub>N<sub>4</sub>, Ag<sub>2</sub>S, and CNAgS25 nanocomposites.</p>
Full article ">Figure 2
<p>N<sub>2</sub>-adsorption isotherm of (<b>a</b>) g-C<sub>3</sub>N<sub>4</sub> and (<b>b</b>) CNAgS25.</p>
Full article ">Figure 3
<p>(<b>a</b>) SEM of CNAgS25, (<b>b</b>) mapping of CNAgS25, (<b>c</b>) mapping of C, (<b>d</b>) mapping of (N), (<b>e</b>) mapping of Ag, (<b>f</b>) mapping of S, (<b>g</b>) EDX of CNAgS25.</p>
Full article ">Figure 4
<p>(<b>a</b>) TEM of CNAgS25, (<b>b</b>) HRTEM of CNAgS25 and (<b>c</b>) SAED of CNAgS25.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) TEM of CNAgS25, (<b>b</b>) HRTEM of CNAgS25 and (<b>c</b>) SAED of CNAgS25.</p>
Full article ">Figure 5
<p>(<b>a</b>) DRS of g-C<sub>3</sub>N<sub>4</sub>, Ag<sub>2</sub>S, CNAgS15, and CNAgS25. (<b>b</b>) Tauc plot of g-C<sub>3</sub>N<sub>4</sub>, Ag2S, CNAgS15, and CNAgS25.</p>
Full article ">Figure 6
<p>PL analysis of g-C<sub>3</sub>N<sub>4</sub>, NAgS15, and CNAgS25.</p>
Full article ">Figure 7
<p>The absorption spectrum for photocatalytic degradation of rhodamine B over the surfaces of g-C3N4, CNAg10, CNAg15, and CNAg25.</p>
Full article ">Figure 8
<p>(<b>a</b>) The variations in the amount of RhB removed (%) under dark and light reactions with the illumination time over the surfaces of g-C<sub>3</sub>N<sub>4</sub>, CNAg10, CNAg15, and CNAg25. (<b>b</b>) The kinetic first-order plot for photocatalytic degradation of RhB dye over the surfaces of g-C<sub>3</sub>N<sub>4</sub>, CNAg10, CNAg15, and CNAg25.</p>
Full article ">Figure 9
<p>Photocatalytic degradation of rhodamine B (2 × 10<sup>−5</sup> M) over CNAgS25 nanocomposite in the presence of 2 × 10<sup>−5</sup> M of the following scavengers: (<b>a</b>) benzoquinone, (<b>b</b>) ammonium oxalate, and (<b>c</b>) isopropanol. (<b>d</b>) PL spectrum of terephthalic acid 2 × 10<sup>−4</sup> M over CNAgS25 nanocomposite at 325 nm excitation wavelength.</p>
Full article ">Figure 9 Cont.
<p>Photocatalytic degradation of rhodamine B (2 × 10<sup>−5</sup> M) over CNAgS25 nanocomposite in the presence of 2 × 10<sup>−5</sup> M of the following scavengers: (<b>a</b>) benzoquinone, (<b>b</b>) ammonium oxalate, and (<b>c</b>) isopropanol. (<b>d</b>) PL spectrum of terephthalic acid 2 × 10<sup>−4</sup> M over CNAgS25 nanocomposite at 325 nm excitation wavelength.</p>
Full article ">Figure 10
<p>Regeneration of CNAgS25 for five consecutive cycles for removal of RhB dye over CNAgS25 nanocomposite.</p>
Full article ">Figure 11
<p>A scheme for electron transportation between g-C<sub>3</sub>N<sub>4</sub> and Ag<sub>2</sub>S semiconductors.</p>
Full article ">Figure 12
<p>Scheme for synthesis of (<b>a</b>) g-C<sub>3</sub>N<sub>4</sub>, (<b>b</b>) Ag<sub>2</sub>S and (<b>c</b>) Ag<sub>2</sub>S/g-C<sub>3</sub>N<sub>4</sub> heterojunction.</p>
Full article ">
24 pages, 3624 KiB  
Review
Recent Advances in the Adsorption of Different Pollutants from Wastewater Using Carbon-Based and Metal-Oxide Nanoparticles
by Shahabaldin Rezania, Negisa Darajeh, Parveen Fatemeh Rupani, Amin Mojiri, Hesam Kamyab and Mohsen Taghavijeloudar
Appl. Sci. 2024, 14(24), 11492; https://doi.org/10.3390/app142411492 - 10 Dec 2024
Viewed by 527
Abstract
In recent years, nanomaterials have gained special attention for removing contaminants from wastewater. Nanoparticles (NPs), such as carbon-based materials and metal oxides, exhibit exceptional adsorption capacity and antimicrobial properties for wastewater treatment. Their unique properties, including reactivity, high surface area, and tunable surface [...] Read more.
In recent years, nanomaterials have gained special attention for removing contaminants from wastewater. Nanoparticles (NPs), such as carbon-based materials and metal oxides, exhibit exceptional adsorption capacity and antimicrobial properties for wastewater treatment. Their unique properties, including reactivity, high surface area, and tunable surface functionalities, make them highly effective adsorbents. They can remove contaminants such as organics, inorganics, pharmaceuticals, medicine, and dyes by adsorption mechanisms. In this review, the effectiveness of different types of carbon-based NPs, including carbon nanotubes (CNTs), graphene-based nanoparticles (GNPs), carbon quantum dots (CQDs), carbon nanofibers (CNFs), and carbon nanospheres (CNSs), and metal oxides, including copper oxide (CuO), zinc oxide (ZnO), iron oxide (Fe2O3), titanium oxide (TiO2), and silver oxide (Ag2O), in the removal of different contaminants from wastewater has been comprehensively evaluated. In addition, their synthesis methods, such as physical, chemical, and biological, have been described. Based on the findings, CNPs can remove 75 to 90% of pollutants within two hours, while MONPs can remove 60% to 99% of dye in 150 min, except iron oxide NPs. For future studies, the integration of NPs into existing treatment systems and the development of novel nanomaterials are recommended. Hence, the potential of NPs is promising, but challenges related to their environmental impact and their toxicity must be considered. Full article
(This article belongs to the Special Issue Water Treatment: From Membrane Processes to Renewable Energies)
Show Figures

Figure 1

Figure 1
<p>Different types of carbon and metal-oxide nanoparticles.</p>
Full article ">Figure 2
<p>Schematic of different adsorption mechanisms.</p>
Full article ">Figure 3
<p>Synthesis methods of NPs.</p>
Full article ">Figure 4
<p>Structural representation of (<b>a</b>) SWCNTs and (<b>b</b>) MWCNTs. Source: Adapted from [<a href="#B57-applsci-14-11492" class="html-bibr">57</a>].</p>
Full article ">Figure 5
<p>Different types of graphene nanomaterials: (<b>a</b>) graphene, (<b>b</b>) GO, (<b>c</b>) rGO, and (<b>d</b>) GQD. Source: [<a href="#B66-applsci-14-11492" class="html-bibr">66</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>c</b>) Formation of a cup-stacked CNF structure and a (<b>d</b>) platelet CNF structure. Source: adapted from [<a href="#B84-applsci-14-11492" class="html-bibr">84</a>].</p>
Full article ">Figure 7
<p>Schematic of the adsorption mechanism using magnetic nanosheets. Source: [<a href="#B121-applsci-14-11492" class="html-bibr">121</a>].</p>
Full article ">Figure 8
<p>Different crystalline structures of TiO<sub>2</sub> nanomaterials: anatase, rutile, and brookite. The red ball represents the Ti<sup>2+</sup> ion, and the white ball is O<sub>2</sub><sup>−</sup>. Source: [<a href="#B154-applsci-14-11492" class="html-bibr">154</a>].</p>
Full article ">Figure 9
<p>Crystal structures of hematite, magnetite, and maghemite. Source: [<a href="#B164-applsci-14-11492" class="html-bibr">164</a>].</p>
Full article ">Figure 10
<p>Different cost allocations of nanomaterials in the wastewater treatment process.</p>
Full article ">
14 pages, 2844 KiB  
Article
Green and Eco-Friendly Egg White–TiO2 Hydrogel with Enhanced Antimicrobial, Adsorptive, and Photocatalytic Properties
by Mei Zhang and Xu Wang
Catalysts 2024, 14(12), 899; https://doi.org/10.3390/catal14120899 - 8 Dec 2024
Viewed by 669
Abstract
The design of multi-purpose decontaminants with environmentally friendly characteristics, low cost, and high efficiency in removing pollutants from the environment is an effective and economic strategy for maintaining the long-term development of the ecosystem. Based on the strategy of killing two birds with [...] Read more.
The design of multi-purpose decontaminants with environmentally friendly characteristics, low cost, and high efficiency in removing pollutants from the environment is an effective and economic strategy for maintaining the long-term development of the ecosystem. Based on the strategy of killing two birds with one stone, an egg white (EW)/TiO2 hydrogel with a porous structure is devised as a bio-adsorbent using waste eggs nearing their expiration date for simultaneously achieving the efficient removal of organic dyes and the inactivation of microorganisms from industrial wastewater. The characterizations of its morphology and composition using scanning electron microscopy (SEM), the Brunauer–Emmett–Teller (BET) theory, energy-dispersive spectrometry (EDS), Fourier transform infrared spectroscopy (FTIR), and a thermogravimetric analyzer (TGA) validate the successful synthesis of EW/TiO2. The maximum adsorption capacity of EW/TiO2 is 333.172 mg∙mL−1 according to the Langmuir model. The photodegradation of a methyl blue (MB) solution under irradiation via a xenon lamp is used to assess the photocatalytic behavior of EW/TiO2. Among the different samples, the 5 wt% TiO2-doped EW/TiO2 hydrogel shows an efficiency of 99% for 120 min of irradiation. Finally, the antibacterial properties of the EW/TiO2 hydrogel are evaluated by calculating its bacterial survival rate against Escherichia coli (E. coli). The EW/TiO2 photocatalyst exhibits a photocatalytic inactivation efficiency of 90.4%, indicating that the EW/TiO2 hydrogel possesses positive antibacterial activity via effectively inhibiting the growth of the bacteria, which is suitable for industrial wastewater treatment over a long period of time. Full article
(This article belongs to the Special Issue Photocatalytic Nanomaterials for Environmental Purification)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The morphology analysis and elemental characterization of the EW/TiO<sub>2</sub> hydrogel. SEM images of (<b>a</b>) EW hydrogel, (<b>b</b>) EW/TiO<sub>2</sub> (2.5%), (<b>c</b>) EW/TiO<sub>2</sub> (5%), (<b>d</b>) EW/TiO<sub>2</sub> (7.5%), and (<b>e</b>) EW/TiO<sub>2</sub> (10%). The addition of TiO<sub>2</sub> into the EW hydrogel makes the surface rougher, providing more active sites. (<b>f</b>) EDS of the EW and EW/TiO<sub>2</sub> hydrogels. Elemental characterizations verify the successful synthesis of the EW/TiO<sub>2</sub> hydrogel.</p>
Full article ">Figure 2
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isotherms of the EW/TiO<sub>2</sub> hydrogel. The EW/TiO<sub>2</sub> hydrogel with a specific surface area of 30.41 m<sup>2</sup>∙g<sup>−1</sup> exhibits a type III isotherm and visible H2 hysteresis loop in the isotherm. (<b>b</b>) The pore size distribution of the EW/TiO<sub>2</sub> hydrogel. The hydrogel with a pore size from 20 to 50 nm displays mesopores.</p>
Full article ">Figure 3
<p>(<b>a</b>) FTIR spectra of the EW/TiO<sub>2</sub> hydrogel certify the successful synthesis of the EW/TiO<sub>2</sub> hydrogel. (<b>b</b>) TGA analysis confirms that the fabricated EW/TiO<sub>2</sub> possesses favorable thermal stability.</p>
Full article ">Figure 4
<p>MB removal efficiency of different adsorbents for 60 min at 25 °C. The optimal dosage of TiO<sub>2</sub> in EW/TiO<sub>2</sub> is 5%. Error bars are based on the average of five measurements.</p>
Full article ">Figure 5
<p>Effect of adsorbent dosage on the adsorption of MB by the EW/TiO<sub>2</sub> hydrogel at 25 °C. Error bars are based on the average of five measurements. The optimal EW/TiO<sub>2</sub> hydrogel dosage is selected as 10 mg.</p>
Full article ">Figure 6
<p>(<b>a</b>) Adsorption kinetics of the EW/TiO<sub>2</sub> hydrogel for MB. The pseudo-first-order model is more suitable for describing the adsorption kinetics of MB on EW/TiO<sub>2</sub>. (<b>b</b>) The intra-particle diffusion model of the EW/TiO<sub>2</sub> hydrogel for MB. The adsorption process contains two steps: (I) MB in the solution diffuses to the surface of the EW/TiO<sub>2</sub> hydrogel; (II) the interaction between the EW/TiO<sub>2</sub> hydrogel and MB.</p>
Full article ">Figure 7
<p>Adsorption isotherm of the EW/TiO<sub>2</sub> hydrogel for MB. The adsorption of MB on the EW/TiO<sub>2</sub> hydrogel is well-fitted with the Langmuir isotherm model.</p>
Full article ">Figure 8
<p>Photocatalytic behavior of MB over EW/TiO<sub>2</sub> under an irradiation of a 300 W xenon lamp. Error bars are based on the average of five measurements.</p>
Full article ">Figure 9
<p>(<b>a</b>) OD<sub>600</sub> of bacteria showing the distinctively different growth of bacteria in suspensions with different samples. (<b>b</b>) Comparison of the survival rate of bacteria in suspensions with different samples. Error bars are based on the average of five measurements.</p>
Full article ">Figure 10
<p>Viability of HEK 293 cells incubated for 24 h in the presence of the EW/TiO<sub>2</sub> hydrogel (5~25 mg) using the MTT assay. Error bars are based on the average of five measurements.</p>
Full article ">
18 pages, 3212 KiB  
Article
Sustainable Green Synthesis of ZnO Nanoparticles from Bromelia pinguin L.: Photocatalytic Properties and Their Contribution to Urban Habitability
by Manuel de Jesus Chinchillas-Chinchillas, Horacio Edgardo Garrafa Galvez, Victor Manuel Orozco Carmona, Hugo Galindo Flores, Jose Belisario Leyva Morales, Mizael Luque Morales, Mariel Organista Camacho and Priscy Alfredo Luque Morales
Sustainability 2024, 16(23), 10745; https://doi.org/10.3390/su162310745 - 7 Dec 2024
Viewed by 640
Abstract
Aguama (Bromelia pinguin L.), a plant belonging to the Bromeliaceae family, possesses a rich content of organic compounds historically employed in traditional medicine. This research focuses on the sustainable synthesis of ZnO nanoparticles via an eco-friendly route using 1, 2, and 4% [...] Read more.
Aguama (Bromelia pinguin L.), a plant belonging to the Bromeliaceae family, possesses a rich content of organic compounds historically employed in traditional medicine. This research focuses on the sustainable synthesis of ZnO nanoparticles via an eco-friendly route using 1, 2, and 4% of Aguama peel extract. This method contributes to environmental sustainability by reducing the use of hazardous chemicals in nanoparticle production. The optical properties, including the band gap, were determined using the TAUC model through Ultraviolet–Visible Spectroscopy (UV–Vis). The photocatalytic activity was evaluated using three widely studied organic dyes (methylene blue, methyl orange, and rhodamine B) under both solar and UV radiation. The results demonstrated that the ZnO nanoparticles, characterized by a wurtzite-type crystalline structure and particle sizes ranging from 68 to 76 nm, exhibited high thermal stability and band gap values between 2.60 and 2.91 eV. These nanoparticles successfully degraded the dyes completely, with methylene blue degrading in 40 min, methyl orange in 70 min, and rhodamine B in 90 min. This study underscores the potential of Bromelia pinguin L. extract in advancing sustainable nanoparticle synthesis and its application in environmental remediation through efficient photocatalysis. Full article
Show Figures

Figure 1

Figure 1
<p>ATR-IR analysis of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">Figure 2
<p>Morphological analysis of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>,<b>b</b>) Micrographs of BP 1%-ZnO, (<b>c</b>) size distribution of BP 1%-ZnO, (<b>d</b>,<b>e</b>) Micrographs of BP 2%-ZnO, (<b>f</b>) size distribution of BP 2%-ZnO, (<b>g</b>,<b>h</b>) Micrographs of BP 4%-ZnO and (<b>i</b>) size distribution of BP 4%-ZnO.</p>
Full article ">Figure 3
<p>XRD spectra of the nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>) BP 1%-ZnO nanoparticles, (<b>b</b>) BP 2%-ZnO nanoparticles and (<b>c</b>) BP 4%-ZnO nanoparticles.</p>
Full article ">Figure 4
<p>TGA/DSC results of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">Figure 5
<p>BET analysis of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">Figure 6
<p>UV–Vis spectra of nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">Figure 7
<p>Band gaps of the nanoparticles synthesized using 1%, 2%, and 4% of <span class="html-italic">Bromelia pinguin</span> L. calculated with the help of the TAUC model.</p>
Full article ">Figure 8
<p>Formation mechanism of nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">Figure 9
<p>Photocatalytic activity of nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L. (<b>a</b>) Degradation of MB under solar radiation; (<b>b</b>) degradation of MB under UV radiation; (<b>c</b>) MO degradation under solar radiation; (<b>d</b>) degradation of MO under UV radiation; (<b>e</b>) degradation of RhB under solar radiation; and (<b>f</b>) degradation of RhB under UV radiation.</p>
Full article ">Figure 10
<p>Proposed photocatalytic degradation mechanism of ZnO nanoparticles biosynthesized with <span class="html-italic">Bromelia pinguin</span> L.</p>
Full article ">
20 pages, 1250 KiB  
Review
Current Trends in Development and Use of Polymeric Ion-Exchange Resins in Wastewater Treatment
by Nicoleta Mirela Marin, Mihai Nita Lazar, Marcela Popa, Toma Galaon and Luoana Florentina Pascu
Materials 2024, 17(23), 5994; https://doi.org/10.3390/ma17235994 - 6 Dec 2024
Viewed by 506
Abstract
Drinking and wastewater are to be treated for safe human consumption and for keeping surface waters clean. There are multiple water purification procedures, but the use of ion-exchange resins significantly enhances water purification efficiency. This review was targeted on highlighting the concept and [...] Read more.
Drinking and wastewater are to be treated for safe human consumption and for keeping surface waters clean. There are multiple water purification procedures, but the use of ion-exchange resins significantly enhances water purification efficiency. This review was targeted on highlighting the concept and classification of polymeric ion-exchange resins as well as pointing out their real-world applications. Their successful use for purification purposes has been linked to their chemical structure, simplicity of operation, accessibility, and reusability. Therefore, polymeric ion-exchange resins have been used for the removal of a wide range of organic and inorganic pollutants such as pharmaceutical compounds, dyes, organic matter, metals, and many others. Ion-exchange resins are obtained directly by synthesis methods or grafting ionizable groups on polymer matrix in order to ensure continuous improvement. Furthermore, the newly designed ion-exchange resins take into consideration biodegradability potential towards obtaining ecofriendly compounds. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structure of a cation-exchange resin with sulfonic groups (<b>a</b>) and anion-exchange resin with quaternary amine groups (<b>b</b>) (adapted according to reference [<a href="#B60-materials-17-05994" class="html-bibr">60</a>]).</p>
Full article ">Figure 2
<p>Synthesis of a strongly acidic cation exchanger based on styrene–divinylbenzene (adapted according to reference [<a href="#B62-materials-17-05994" class="html-bibr">62</a>]).</p>
Full article ">Figure 3
<p>Synthesis of the polymer support for a weakly acidic cation-exchange resin (adapted according to reference [<a href="#B62-materials-17-05994" class="html-bibr">62</a>]).</p>
Full article ">Figure 4
<p>The reaction to obtain a strongly basic anion-exchange resin (adapted according to reference [<a href="#B62-materials-17-05994" class="html-bibr">62</a>]).</p>
Full article ">Figure 5
<p>Classification of ion-exchange resins (adapted according to reference [<a href="#B63-materials-17-05994" class="html-bibr">63</a>]).</p>
Full article ">
20 pages, 5748 KiB  
Review
Nanolabels Prepared by the Entrapment or Self-Assembly of Signaling Molecules for Colorimetric and Fluorescent Immunoassays
by Ning Xia, Yadi Li, Cancan He and Dehua Deng
Biosensors 2024, 14(12), 597; https://doi.org/10.3390/bios14120597 - 6 Dec 2024
Viewed by 460
Abstract
Nanomaterials have attracted significant attention as signal reporters for immunoassays. They can directly generate detectable signals or release a large number of signaling elements for readout. Among various nanolabels, nanomaterials composed of multiple signaling molecules have shown great potential in immunoassays. Generally, signaling [...] Read more.
Nanomaterials have attracted significant attention as signal reporters for immunoassays. They can directly generate detectable signals or release a large number of signaling elements for readout. Among various nanolabels, nanomaterials composed of multiple signaling molecules have shown great potential in immunoassays. Generally, signaling molecules can be entrapped in nanocontainers or self-assemble into nanostructures for signal amplification. In this review, we summarize the advances of signaling molecules-entrapped or assembled nanomaterials for colorimetric and fluorescence immunoassays. The nanocontainers cover liposomes, polymers, mesoporous silica, metal–organic frameworks (MOFs), various nanosheets, nanoflowers or nanocages, etc. Signaling molecules mainly refer to visible and/or fluorescent organic dyes. The design and application of immunoassays are emphasized from the perspective of nanocontainers, analytes, and analytical performances. In addition, the future challenges and research trends for the preparation of signaling molecules-entrapped or assembled nanolabels are briefly discussed. Full article
(This article belongs to the Special Issue Biosensors Based on Self-Assembly and Boronate Affinity Interaction)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>A</b>) Schematic of the liposome-amplified plasmonic immunoassay (LAPIA). The detection steps include the capture (<b>a</b>) and recognition (<b>b</b>) of the target, attachment of streptavidin (<b>c</b>), coupling of biotin-conjugated cysteine-contained liposomes (<b>d</b>), breakdown of the liposomes (<b>e</b>), and release of cysteine to trigger the aggregation of AuNPs (<b>f</b>) [<a href="#B27-biosensors-14-00597" class="html-bibr">27</a>]. Copyright 2015 American Chemical Society. (<b>B</b>) Schematic illustration of signal-on competitive-type colorimetric immunoassay for the detection of streptomycin (STR) on monoclonal anti-STR antibody-coated microplate using glucose-loaded liposome as the signal tracer labeled with STR-bovine serum albumin (BSA) conjugate: (<b>a</b>) competitive-type immunoreaction and (<b>b</b>) glucose oxidase (GOx)-triggered the change of the Fe(II)-Phen system in the absorbance and visual color by the reaction of the produced H<sub>2</sub>O<sub>2</sub> with iron(II) [<a href="#B41-biosensors-14-00597" class="html-bibr">41</a>]. Copyright 2018 Elsevier.</p>
Full article ">Figure 2
<p>Schematic presentation of the heterogeneous sandwich immunoassay with PSP for loading load C153, hemin, or microperoxidase MP11 based on different signal generation strategies and photo/chemiluminescence detection [<a href="#B51-biosensors-14-00597" class="html-bibr">51</a>]. Copyright 2024 American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>A</b>) (<b>a</b>) Synthesis and derivatization of TP@PEI/Ab<sub>2</sub>-MSNs and (<b>b</b>) steps of the enzyme-free immunosorbent assay of PSA using for amplified colorimetric detection in a 96-well plate [<a href="#B63-biosensors-14-00597" class="html-bibr">63</a>]. Copyright 2018 American Chemical Society. (<b>B</b>) Schematic illustration of the magnetic bead (MB)-based colorimetric immunoassay of PSA by the redox cycling with Ab<sub>2</sub>-MSN-PQQ as the nanolabel [<a href="#B64-biosensors-14-00597" class="html-bibr">64</a>]. Copyright 2019 Elsevier. (<b>C</b>) Schematic illustration of the fluorescence immunoassay based on target-induced competitive displacement reaction between glucose and mannose for Con A accompanying cargo (rhodamine B) release from magnetic mesoporous silica nanoparticles (MMSNs) [<a href="#B65-biosensors-14-00597" class="html-bibr">65</a>]. Copyright 2013 American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>A</b>) (<b>a</b>) Preparation of MOFs NH<sub>2</sub>-MIL-53(Al) and (<b>b</b>) schematic illustration of competitive FIA of AFB1 [<a href="#B81-biosensors-14-00597" class="html-bibr">81</a>]. Copyright 2019 American Chemical Society. (<b>B</b>) Schematic illustration of the synthetic procedure of MILL-88@TcP nanozyme-based detection probe (<b>top</b>) and the procedure of this developed N-ELISA for <span class="html-italic">S. typhimurium</span> detection in milk (<b>bottom</b>) [<a href="#B89-biosensors-14-00597" class="html-bibr">89</a>]. Copyright 2024 Elsevier.</p>
Full article ">Figure 5
<p>(<b>A</b>) Schematic representation of MPDA@TP-linked immunosorbent assay (MLISA) for α-fetoprotein (AFP) on anti-AFP capture antibody (CAb)-modified microplate using anti-AFP detection antibody (DAb)-labeled MPDA@TP with a sandwich-type immunoreaction mode [<a href="#B92-biosensors-14-00597" class="html-bibr">92</a>]. Copyright 2018 American Chemical Society. (<b>B</b>) The synthesis of UiO, UiOL, UiOL@AIEgens, and UiOL@AIEgens-mAbs probe (<b>a</b>), and the UiOL@AIEgens-based POC LFIS for visual and quantitative dual-modal detection of AFB1 (<b>b</b>) [<a href="#B96-biosensors-14-00597" class="html-bibr">96</a>]. Copyright 2024 Elsevier.</p>
Full article ">Figure 6
<p>(<b>A</b>) (<b>a</b>) The preparation process for the signal label of PP-Ab<sub>2</sub>-cC<sub>3</sub>N<sub>4</sub>, and (<b>b</b>) schematic illustration of the PILISA for the detection of CEA in 96-well PS plates [<a href="#B100-biosensors-14-00597" class="html-bibr">100</a>]. Copyright 2017 Elsevier. (<b>B</b>) Schematic of AuNF@Fluorescein@SA preparation, and AuNF@Fluorescein@SA-based dual-mode fluorescent and colorimetric immunoassay [<a href="#B101-biosensors-14-00597" class="html-bibr">101</a>]. Copyright 2018 Elsevier.</p>
Full article ">Figure 7
<p>(<b>A</b>) Synthetic procedure for SAD carriers and their applications in ICAs. (<b>B</b>) Comparison of SAD-ICAs with three modes and a traditional nanomaterial—ICA (take AuNPs as an example)—for the detection of ZEN [<a href="#B102-biosensors-14-00597" class="html-bibr">102</a>]. Copyright 2021 American Chemical Society.</p>
Full article ">Figure 8
<p>(<b>A</b>) Principle of immunoassay using antigen-decorated perylene microparticles [<a href="#B109-biosensors-14-00597" class="html-bibr">109</a>]. Copyright 2000 Elsevier. (<b>B</b>) Principle of a sandwich fluorescent immunoassay using nanocrystalline fluorescein diacetate (FDA) conjugates [<a href="#B103-biosensors-14-00597" class="html-bibr">103</a>]. Copyright 2004 American Chemical Society.</p>
Full article ">Figure 9
<p>(<b>A</b>) Workflow of porphyrin nanoparticle-based signal amplification sandwich assays for the detection of biomolecules [<a href="#B112-biosensors-14-00597" class="html-bibr">112</a>]. Copyright 2016 American Chemical Society. (<b>B</b>) Scheme of sandwich-type TLISA for the detection of IL-6 [<a href="#B115-biosensors-14-00597" class="html-bibr">115</a>]. Copyright 2019 American Chemical Society.</p>
Full article ">Figure 10
<p>(<b>A</b>) Schematic diagram illustrating the principle of using organic nanoparticles as biolabels for immunodipsticks [<a href="#B119-biosensors-14-00597" class="html-bibr">119</a>]. Copyright 2011 Elsevier. (<b>B</b>) Schematic representation of the strategy of integrating an SAN-LFA for the detection of cardiac biomarkers [<a href="#B120-biosensors-14-00597" class="html-bibr">120</a>]. Copyright 2016 American Chemical Society.</p>
Full article ">
20 pages, 12643 KiB  
Article
Titanium Dioxide 1D Nanostructures as Photocatalysts for Degradation and Removal of Pollutants in Water
by Dora María Frías Márquez, José Ángel Méndez González, Rosendo López González, Cinthia García Mendoza, Francisco Javier Tzompantzi Morales, Patricia Quintana Owen and Mayra Angélica Alvarez Lemus
Catalysts 2024, 14(12), 896; https://doi.org/10.3390/catal14120896 - 6 Dec 2024
Viewed by 550
Abstract
The oxidation of organic pollutants in water is the most reported application of a Titanium dioxide (TiO2) photocatalyst. During the last decade, photoreduction with TiO2 has also been explored but simultaneous capabilities for unmodified TiO2 have not been reported [...] Read more.
The oxidation of organic pollutants in water is the most reported application of a Titanium dioxide (TiO2) photocatalyst. During the last decade, photoreduction with TiO2 has also been explored but simultaneous capabilities for unmodified TiO2 have not been reported yet. Here, we reported on the fabrication of TiO2 nanorods using hydrothermal treatment and compared the effect of two different TiO2 powders as the starting material: P-25 and TiO2 sol–gel (N-P25 and N-TiO2, respectively) which were further calcined at 400 °C (N-P25-400 and N-TiO2-400). XPS and XRD analyses confirmed the presence of sodium and hydrogen titanates in N-P25, but also an anatase structure for N-TiO2. The specific surface area of the calcined samples decreased compared to the dried samples. Photocatalytic activity was evaluated using phenol and methyl orange for degradation, whereas 4-nitrophenol was used for photoreduction. Irradiation of the suspension was performed under UV light (λ = 254 nm). The results demonstrated that the nanorods calcined at 400 °C were more photoactive since methyl orange (20 ppm) degradation reached 86% after 2 h, when N-TiO2-400 was used. On the other hand, phenol (20 ppm) was completely degraded by the presence of N-P25-400 after 2 h. Photoreduction of 4-nitrophenol (5 ppm) was achieved by the N-TiO2-400 during the same period. These results demonstrate that the presence of Ti3+ and the source of TiO2 have a significant effect on the photocatalytic activity of TiO2 nanorods. Additionally, the removal of methylene blue (20 ppm) was performed, demonstrating that N-TiO2 exhibited a high adsorption capacity for this dye. Full article
(This article belongs to the Special Issue Advances in Photocatalytic Degradation)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction patterns for (<b>A</b>) the TiO<sub>2</sub> starting powders, (<b>B</b>) the hydrothermally treated TiO2 samples, and (<b>C</b>) the nanostructures after calcination at 400 °C.</p>
Full article ">Figure 2
<p>Raman spectra of (<b>A</b>) P25 and (<b>B</b>) TiO<sub>2</sub> sol–gel-derived nanostructures. Red dashed lines indicate the characteristic signals attributed to B1g and Eg vibration modes for the anatase crystalline phase.</p>
Full article ">Figure 3
<p>DR UV-Visible spectra of (<b>A</b>) the nanomaterials obtained from P25 and (<b>B</b>) the nanomaterials prepared using TiO<sub>2</sub> sol–gel.</p>
Full article ">Figure 4
<p>Highly resolved X-ray photoelectron spectra for the nanomaterials prepared. (<b>A</b>,<b>B</b>) correspond with P25, and (<b>C</b>,<b>D</b>) correspond with TiO<sub>2</sub> sol–gel-derived materials.</p>
Full article ">Figure 5
<p>Field Emission Scanning Electron Microscopies of the samples. Yellow lines indicate representative measured particles.</p>
Full article ">Figure 6
<p>(<b>A</b>,<b>B</b>) display the N<sub>2</sub> adsorption–desorption isotherms for the P25 and TiO<sub>2</sub> sol–gel-derived photocatalysts, respectively. The pore size distribution obtained from the desorption branch and the BJH method for (<b>C</b>) P25 and (<b>D</b>) TiO<sub>2</sub> sol–gel samples.</p>
Full article ">Figure 7
<p>(<b>A</b>) Kinetic profiles for the degradation of the 20 ppm methyl orange solution adjusting to first-order kinetics, and (<b>B</b>) kinetic profiles for the degradation of the 20 ppm phenol solution adjusted to zero-order kinetics under UV light (λ = 254 nm); (<b>C</b>) the adsorption capacity of N-TiO<sub>2</sub> for different concentrations of methylene blue (MB) after 60 min in the dark; the (<b>D</b>) photoreduction of 4-Nitrophenol with the N-TiO<sub>2</sub>-400 photocatalyst under UV light (λ = 254 nm).</p>
Full article ">Figure 8
<p>HRTEM images for the (<b>A</b>,<b>B</b>) N-TiO<sub>2</sub> and (<b>D</b>,<b>E</b>) N-TiO<sub>2</sub>-400 photocatalysts (<b>C</b>,<b>F</b>) represent the selected area electron diffraction (SAED) patterns of the N-TiO<sub>2</sub> and N-TiO<sub>2</sub>-400, respectively.</p>
Full article ">
15 pages, 4874 KiB  
Article
Efficient Photocatalytic Degradation of Methylene Blue and Methyl Orange Using Calcium-Polyoxometalate Under Ultraviolet Irradiation
by Suhair A. Bani-Atta, A. A. A. Darwish, Leena Shwashreh, Fatimah A. Alotaibi, Jozaa N. Al-Tweher, Hatem A. Al-Aoh and E. F. M. El-Zaidia
Processes 2024, 12(12), 2769; https://doi.org/10.3390/pr12122769 - 5 Dec 2024
Viewed by 492
Abstract
With the increasing demand for eco-friendly water treatment solutions, the development of novel photocatalysts such as calcium polyanion (Ca-POM) plays a vital role in mitigating industrial wastewater pollution. In this research, calcium polyanion, H60N6Na2Ca2W12 [...] Read more.
With the increasing demand for eco-friendly water treatment solutions, the development of novel photocatalysts such as calcium polyanion (Ca-POM) plays a vital role in mitigating industrial wastewater pollution. In this research, calcium polyanion, H60N6Na2Ca2W12O60 (Ca–POM), was successfully synthesized via a self-assembly reaction from metal-oxide subunits. The synthesized Ca–POM was verified to have a polycrystalline structure with a broad size distribution, with an average particle diameter of approximately 623.62 nm. Powder X-ray diffraction (XRD) analysis confirmed the polycrystalline structure of the Ca–POM, with a calculated band gap energy of 3.29 eV. The photocatalytic behavior of the Ca-POM sample was tested with two model dyes, methylene blue (MB) and methyl orange (MO). The reaction mixture was then exposed to ultraviolet (UV) irradiation for durations ranging from 20 to 140 min. The synthesized cluster demonstrated photocatalytic efficiency (PCE%) values of 81.21% for MB and 25.80% for MO. This work offers a valuable basis for applying Ca–POM as a heterogeneous photocatalyst for treating industrial wastewater organic pollutants and highlights the potential of Ca–POM in sustainable water treatment applications. Full article
Show Figures

Figure 1

Figure 1
<p>A spectrum of FT-IR for Ca–POM.</p>
Full article ">Figure 2
<p>XRD pattern of Ca-POM.</p>
Full article ">Figure 3
<p>Thermogravimetric analysis (TGA) of Ca–POM.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM image and (<b>b</b>) particle size distribution of Ca–POM.</p>
Full article ">Figure 5
<p>UV-Vis absorbance spectrum of Ca–POM solution.</p>
Full article ">Figure 6
<p>Relation between (<span class="html-italic">αhν</span>)<sup>1/2</sup> and <span class="html-italic">hν</span> of Ca–POM.</p>
Full article ">Figure 7
<p>Absorbance of methylene blue (MB) solution at different UV irradiation times: (<b>a</b>) without and (<b>b</b>) with Ca-POM catalyst.</p>
Full article ">Figure 7 Cont.
<p>Absorbance of methylene blue (MB) solution at different UV irradiation times: (<b>a</b>) without and (<b>b</b>) with Ca-POM catalyst.</p>
Full article ">Figure 8
<p>Absorbance of methyl orange (MO) solution at different UV irradiation times: (<b>a</b>) without and (<b>b</b>) with Ca–POM catalyst.</p>
Full article ">Figure 8 Cont.
<p>Absorbance of methyl orange (MO) solution at different UV irradiation times: (<b>a</b>) without and (<b>b</b>) with Ca–POM catalyst.</p>
Full article ">Figure 9
<p>Photocatalytic efficiency (PCE%) of Ca–POM in degrading (<b>a</b>) MB and (<b>b</b>) MO under UV light.</p>
Full article ">Figure 10
<p>Relationship between ln(Ct/Co) and irradiation time for the degradation of (<b>a</b>) MB and (<b>b</b>) MO with the Ca-POM catalyst.</p>
Full article ">Scheme 1
<p>The proposed photocatalytic mechanisms involved in the simultaneous oxidation of dyes using the Ca–POM catalyst.</p>
Full article ">
20 pages, 7890 KiB  
Article
Insights into Novel Doping Effect of Fe-Doped ZnS Nanostructures Derived from Oxystelma esculentum: Kinetics-Based Photocatalysis, Nitrogen Fixation, and Antifungal Efficacy
by Mohammad Ehtisham Khan
Catalysts 2024, 14(12), 888; https://doi.org/10.3390/catal14120888 - 4 Dec 2024
Viewed by 551
Abstract
Implementing greener approaches is a sustainable and eco-friendly methodology for nanocomposite synthesis. This work reports the sustainable fabrication of Fe-doped ZnS (Fe0.3Zn0.7S) nanocomposite and its broad-spectrum applications. The systematic characterization was carried out using several advanced analytical techniques. DLS, [...] Read more.
Implementing greener approaches is a sustainable and eco-friendly methodology for nanocomposite synthesis. This work reports the sustainable fabrication of Fe-doped ZnS (Fe0.3Zn0.7S) nanocomposite and its broad-spectrum applications. The systematic characterization was carried out using several advanced analytical techniques. DLS, Zeta potential, SEM, XPS, and TEM performed morphological and size assessments of the engineered nanocomposite. Eventually, XRD provided valuable insights into the crystalline behavior of nanocomposite. The nanocomposites were then treated against the organic dye Safranin O, which displayed 93% degradation within an hour with the rate constant value of 0.0326 min−1. Parameters influencing the percentage degradation, such as temperature, pH, etc., were also discussed. Moreover, an LCMS test was also conducted to evaluate the presence of reactive intermediates. Safranin O’s degradation was confirmed by identifying intermediate products, such as compounds with m/z values of 335.84, 321.81, 306.79, 292.77, and 257.32, which were indicative of progressive dye breakdown. Finally, the photocatalytic enactment examination verified that the prepared nanocomposite’s nitrogen fixation rate (38.96 µmolg−1) was way greater (~4 times) than the pristine compound. In addition, prepared nanoparticles demonstrated a befitting ability to eliminate a wide range of threatening pathogenic fungi. The doping of Fe into ZnS further enhanced the inhibition against Fusarium oxysporum. Full article
(This article belongs to the Special Issue Recent Advances in Photocatalysis Research in Asia)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>–<b>F</b>) TEM images of synthesized Fe-doped ZnS nanocomposites indicating the morphological illustration of the prepared nanocomposites at different scales.</p>
Full article ">Figure 2
<p>(<b>A</b>–<b>F</b>) SEM images of the prepared Fe-doped ZnS nanocomposite indicate the morphology of the engineered nanocomposite at different scales.</p>
Full article ">Figure 3
<p>(<b>A</b>) FTIR spectrum of the prepared Fe-doped ZnS nanocomposite, and (<b>B</b>) XRD spectrum of the Fe-doped ZnS nanocomposite.</p>
Full article ">Figure 4
<p>(<b>A</b>) Survey spectrum of Fe-doped ZnS nanocomposite, (<b>B</b>) the high-resolution spectrum of Fe, and (<b>C</b>) the high-resolution spectrum of Zn, and (<b>D</b>) high-resolution XPS spectrum of Sulphur.</p>
Full article ">Figure 5
<p>(<b>A</b>) Pure UV-Vis spectrum of Safranin O, (<b>B</b>) UV-Vis spectra of Safranin O after the addition of NaBH<sub>4</sub>, (<b>C</b>) time-dependent spectra of degradation of Safranin O, and (<b>D</b>) kinetically plotted graph of photocatalytic activity of SO.</p>
Full article ">Figure 6
<p>(<b>A</b>) The effect of different pH values on the percentage degradation of SO, (<b>B</b>) a graphical illustration of the effect of photocatalyst dosage on percentage degradation, (<b>C</b>) the effect of dye concentration on the percentage degradation of SO, and (<b>D</b>) a scavenging test for the photocatalytic performance of SO.</p>
Full article ">Figure 7
<p>(<b>A</b>) Evaluation of NH<sub>3</sub> by the implementation of UV-vis light, (<b>B</b>,<b>C</b>) photocatalytic nitrogen fixation rates under Ar and dark, and (<b>D</b>) cycling tests for Fe-ZnS under UV-vis light.</p>
Full article ">Figure 8
<p>Fungal growth inhibition. (<b>A</b>) Control. (<b>B</b>) Treatment with plant extract. (<b>C</b>) Treatment with pristine compound. (<b>D</b>) Action of Fe-ZnS toward fungal growth inhibition.</p>
Full article ">Scheme 1
<p>Schematic mechanism scheme indicating the photocatalysis and nitrogen fixation using Fe-doped ZnS nanocomposites.</p>
Full article ">
19 pages, 4678 KiB  
Article
Ionic Crosslinking of Linear Polyethyleneimine Hydrogels with Tripolyphosphate
by Luis M. Araque, Antonia Infantes-Molina, Enrique Rodríguez-Castellón, Yamila Garro-Linck, Belén Franzoni, Claudio J. Pérez, Guillermo J. Copello and Juan M. Lázaro-Martínez
Gels 2024, 10(12), 790; https://doi.org/10.3390/gels10120790 - 3 Dec 2024
Viewed by 649
Abstract
In this work, the mechanical properties of hydrogels based on linear polyethyleneimine (PEI) chemically crosslinked with ethyleneglycoldiglycidyl ether (EGDE) were improved by the ionic crosslinking with sodium tripolyphosphate (TPP). To this end, the quaternization of the nitrogen atoms present in the PEI structure [...] Read more.
In this work, the mechanical properties of hydrogels based on linear polyethyleneimine (PEI) chemically crosslinked with ethyleneglycoldiglycidyl ether (EGDE) were improved by the ionic crosslinking with sodium tripolyphosphate (TPP). To this end, the quaternization of the nitrogen atoms present in the PEI structure was conducted to render a network with a permanent positive charge to interact with the negative charges of TPP. The co-crosslinking process was studied by 1H high-resolution magic angle spinning (1H HRMAS) NMR and X-ray photoelectron spectroscopy (XPS) in combination with organic elemental analysis and inductively coupled plasma mass spectrometry (ICP-MS). In addition, the mobility and confinement of water molecules within the co-crosslinked hydrogels were studied by low-field 1H NMR. The addition of small amounts of TPP, 0.03 to 0.26 mmoles of TPP per gram of material, to the PEI-EGDE hydrogel resulted in an increase in the deformation resistance from 320 to 1080%, respectively. Moreover, the adsorption capacity of the hydrogels towards various emerging contaminants remained high after the TPP crosslinking, with maximum loading capacities (qmax) of 77, 512, and 55 mg g−1 at pH = 4 for penicillin V (antibiotic), methyl orange (azo-dye) and copper(II) ions (metal ion), respectively. A significant decrease in the adsorption capacity was observed at pH = 7 or 10, with qmax of 356 or 64 and 23 or 0.8 mg g−1 for methyl orange and penicillin V, respectively. Full article
(This article belongs to the Special Issue Functionalized Gels for Environmental Applications (2nd Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>ATR-FTIR spectra for TPP, PEI-EGDE hydrogel, quaternized PEI-EGDE hydrogel, and P1T0.01, P1T0.05, and P1T0.1 co-crosslinked hydrogels.</p>
Full article ">Figure 2
<p>C 1<span class="html-italic">s</span>, O 1<span class="html-italic">s</span>, N 1<span class="html-italic">s,</span> and P 2<span class="html-italic">p</span> high-resolution XPS spectra for the PEI-EGE hydrogel, and P1T0.01, P1T0.05, and P1T0.1 co-crosslinked hydrogels.</p>
Full article ">Figure 3
<p>Full (<b>A</b>) and partial magnification (<b>B</b>) of the <sup>1</sup>H HRMAS NMR for the PEI-EGDE, quaternized PEI-EGDE, P1T0.01, P1T0.05, and P1T0.1 hydrogels swelled in D<sub>2</sub>O (MAS rate = 4 kHz).</p>
Full article ">Figure 4
<p><sup>31</sup>P direct-polarization <span class="html-italic">ss</span>-NMR spectrum for P1T0.05 hydrogel (MAS rate = 15 kHz). Rotational bands are indicated with an asterisk.</p>
Full article ">Figure 5
<p>T<sub>1</sub>–T<sub>2</sub> maps (<b>A</b>) and T<sub>2</sub> projections for the indicated hydrogels (<b>B</b>).</p>
Full article ">Figure 6
<p>Swelling capacities (<b>A</b>) and TGA curves (<b>B</b>) for the PEI-EGDE, quaternized PEI-EGDE, P1T0.01, P1T0.05, and P1T0.1 hydrogels.</p>
Full article ">Figure 7
<p>Rheological properties of PEI-EGDE, P1T0.01, P1T0.05, and P1T0.1 hydrogels: storage and loss moduli (<b>A</b>) and complex viscosity (<b>B</b>).</p>
Full article ">Figure 8
<p>Adsorption kinetics of MO (<b>A</b>), Cu<sup>2+</sup> ions (<b>B</b>) and PEN (<b>C</b>) by the PEI-EGDE and P1T0.01 hydrogels together with the kinetic model fittings.</p>
Full article ">Figure 9
<p>Adsorption isotherms of MO (<b>A</b>), Cu<sup>2+</sup> ions (<b>B</b>) and PEN (<b>C</b>) by the PEI-EGDE and P1T0.01 hydrogels together with the fittings to the Langmuir model.</p>
Full article ">Scheme 1
<p>Synthetic pathways for the co-crosslinking of PEI-EGDE hydrogels (PEI-EGDE-TPP).</p>
Full article ">
27 pages, 2052 KiB  
Review
Photocatalytic Composites Based on Biochar for Antibiotic and Dye Removal in Water Treatment
by Amra Bratovčić and Vesna Tomašić
Processes 2024, 12(12), 2746; https://doi.org/10.3390/pr12122746 - 3 Dec 2024
Viewed by 917
Abstract
Many semiconductor photocatalysts are characterized by high photostability and non-toxicity but suffer from the limited excitation in the UV part of the spectrum and the fast recombination of the photogenerated electron–hole pairs. To improve the above properties, biochar-supported composite photocatalysts have recently attracted [...] Read more.
Many semiconductor photocatalysts are characterized by high photostability and non-toxicity but suffer from the limited excitation in the UV part of the spectrum and the fast recombination of the photogenerated electron–hole pairs. To improve the above properties, biochar-supported composite photocatalysts have recently attracted much attention. Compared with the pure photocatalyst, the biochar-enriched catalyst has superior specific surface area and high porosity, catalytic efficiency, stability, and recoverability. Biochar can be obtained from various carbon-rich plant or animal wastes by different thermochemical processes such as pyrolysis, hydrothermal carbonization, torrefaction, and gasification. The main features of biochar are its low price, non-toxicity, and the large number of surface functional groups. This paper systematically presents the latest research results on the method of preparation of various composites in terms of the choice of photoactive species and the source of biomass, their physico-chemical properties, the mechanism of the photocatalytic activity, and degradation efficiency in the treatment of organic contaminants (dyes and antibiotics) in an aquatic environment. Particular emphasis is placed on understanding the role of biochar in improving the photocatalytic activity of photoactive species. Full article
(This article belongs to the Special Issue Treatment and Remediation of Organic and Inorganic Pollutants)
Show Figures

Figure 1

Figure 1
<p>Classification of biomass sources based on the initial moisture content and categorization of thermochemical processes based on the aggregate state of the target product.</p>
Full article ">Figure 2
<p>Advantages of biochar in photocatalysis.</p>
Full article ">Figure 3
<p>Schematic illustration of the catalytic degradation of organic pollutant.</p>
Full article ">
Back to TopTop