[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (756)

Search Parameters:
Keywords = methyl orange

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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 586
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, 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 517
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 466
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 ">
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 613
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 ">
14 pages, 16525 KiB  
Article
Preparation and Piezocatalytic Performance of γ-AlON Particles for Dye-Pollutant Degradation Under Ultrasonic Vibration
by Dan Zhu, Yanyan Wang, Le Xiao, Yu Dai and Jian Wu
Molecules 2024, 29(23), 5698; https://doi.org/10.3390/molecules29235698 - 2 Dec 2024
Viewed by 436
Abstract
Piezocatalytic materials have attracted widespread attention in the fields of clean energy and water treatment because of their ability to convert mechanical energy directly into chemical energy. In this study, γ-AlON particles synthesised using carbothermal reduction and nitridation (CRN) were used for the [...] Read more.
Piezocatalytic materials have attracted widespread attention in the fields of clean energy and water treatment because of their ability to convert mechanical energy directly into chemical energy. In this study, γ-AlON particles synthesised using carbothermal reduction and nitridation (CRN) were used for the first time as a novel piezocatalytic material to degrade dye solutions under ultrasonic vibration. The γ-AlON particles exhibited good performance as a piezocatalytic material for the degradation of organic pollutants. After 120 min under ultrasonic vibration, 40 mg portions of γ-AlON particles in 50 mL dye solutions (10 mg/L) achieved 78.06%, 67.74%, 74.29% and 64.62% decomposition rates for rhodamine B (RhB), methyl orange (MO), methylene blue (MB) and crystal violet (CV) solutions, respectively; the fitted k values were 13.35 × 10−3, 10.79 × 10−3, 12.09 × 10−3 and 8.00 × 10−3 min−1, respectively. The piezocatalytic mechanism of γ-AlON particles in the selective degradation of MO was further analysed in free-radical scavenging activity experiments. Hydroxyl radicals (•OH), superoxide radicals (•O2), holes (h+) and electrons (e) were found to be the main active substances in the degradation process. Therefore, γ-AlON particles are an efficient and promising piezocatalytic material for the treatment of dye pollutants. Full article
(This article belongs to the Special Issue Functional Nanomaterials for Energy and Environmental Sustainability)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD and (<b>b</b>) FTIR patterns of γ-AlON particles.</p>
Full article ">Figure 2
<p>(<b>a</b>) XPS spectra of γ-AlON particles and enlarged signals for (<b>b</b>) Al 2<span class="html-italic">p</span>, (<b>c</b>) O 1<span class="html-italic">s</span>, (<b>d</b>) N 1<span class="html-italic">s</span>.</p>
Full article ">Figure 3
<p>SEM images of γ-AlON particles. (<b>a</b>) morphology; (<b>b</b>) morphology of surface step.</p>
Full article ">Figure 4
<p>(<b>a</b>) TEM, (<b>b</b>) HRTEM and (<b>c</b>) SAED and element mapping for (<b>d</b>) Al, (<b>e</b>) O and (<b>f</b>) N of γ-AlON particles.</p>
Full article ">Figure 5
<p>(<b>a</b>) Nitrogen adsorption–desorption isotherm and (<b>b</b>) pore-size distribution curve of γ-AlON particles.</p>
Full article ">Figure 6
<p>(<b>a</b>) Amplitude butterfly curve and (<b>b</b>) phase hysteresis loop of γ-AlON particles.</p>
Full article ">Figure 7
<p>(<b>a</b>) The relative concentration curves, (<b>b</b>) the first-order reaction kinetics curves and (<b>c</b>) the rate constant <span class="html-italic">k</span> of RhB solutions with different dosages of γ-AlON particles under ultrasonic vibration.</p>
Full article ">Figure 8
<p>The cycling runs of degradation of RhB by γ-AlON particles.</p>
Full article ">Figure 9
<p>(<b>a</b>) XRD pattern and (<b>b</b>) SEM image of γ-AlON particles after five cycles of piezocatalytic degradation.</p>
Full article ">Figure 10
<p>The relative concentration curves and the first-order reaction kinetics curves of (<b>a</b>) MO solution, (<b>b</b>) MB solution, (<b>c</b>) CV solution and (<b>d</b>) its rate constant <span class="html-italic">k</span> by γ-AlON particles under ultrasonic vibration.</p>
Full article ">Figure 11
<p>EPR patterns of γ-AlON solutions by DMPO under different ultrasonic vibration times: (<b>a</b>) DMPO-●OH signals and (<b>b</b>) DMPO-●O<sub>2</sub><sup>−</sup> signals.</p>
Full article ">Figure 12
<p>(<b>a</b>) The relative concentration curves. (<b>b</b>) The first-order reaction kinetics curves and (<b>c</b>) the rate constant <span class="html-italic">k</span> of MO solutions by γ-AlON particles under ultrasonic vibration with and without free-radical scavengers.</p>
Full article ">Figure 13
<p>The schematic diagram of piezocatalysis of γ-AlON particles for MO solution.</p>
Full article ">
19 pages, 1899 KiB  
Article
Catalytic Evaluation of an Optimized Heterogeneous Composite Catalyst Derived from Fusion of Tri-Biogenic Residues
by Oyelayo Ajamu Oyedele, Simeon Olatayo Jekayinfa, Abass O. Alade and Christopher Chintua Enweremadu
Biomass 2024, 4(4), 1219-1237; https://doi.org/10.3390/biomass4040068 - 2 Dec 2024
Viewed by 506
Abstract
This study analyzes the elemental and oxide compositions of three selected agricultural residues—Dried Pawpaw Leaves (DPL), Kola Nut Pod (KNP), and Sweet Orange Peel (SOP)—for their potential as heterogeneous catalysts. Energy Dispersive X-ray (EDX) analysis identified calcium (25%) and potassium (29%) as the [...] Read more.
This study analyzes the elemental and oxide compositions of three selected agricultural residues—Dried Pawpaw Leaves (DPL), Kola Nut Pod (KNP), and Sweet Orange Peel (SOP)—for their potential as heterogeneous catalysts. Energy Dispersive X-ray (EDX) analysis identified calcium (25%) and potassium (29%) as the primary elements in DPL and KNP, with calcium oxide (CaO) and potassium oxide (K2O) as the dominant oxides. SOP had a similar composition but lacked vanadium. Calcined residues were analyzed at temperatures ranging from 500 °C to 900 °C using X-ray Fluorescence (XRF), revealing stable silicon dioxide (SiO2) content and temperature-dependent variations in CaO and K2O, indicating their catalytic potential for transesterification processes. Scanning Electron Microscopy (SEM) showed non-uniform, spongy microstructures, enhancing the surface area and catalytic efficiency. Fourier Transform Infrared Spectroscopy (FTIR) identified functional groups essential for catalytic activity, such as hydroxyls, methyl, and carboxyl. X-ray Diffraction (XRD) confirmed the presence of crystalline phases like calcium carbonate and calcium oxide, crucial for catalytic performance. Experimental biodiesel production using a mixture of the calcined residues (33.33% each of KNPA, SOPA, and DPLA) resulted in the highest biodiesel yield at 65.3%. Model summary statistics, including R2 (0.9824) values and standard deviations (0.0026), validated the experimental design, indicating high precision and prediction accuracy. These results suggest that the selected agricultural residues, when calcined and mixed properly, can serve as effective heterogeneous catalysts, with significant implications for biodiesel production, supporting previous research on the importance of calcium in catalytic processes. Full article
Show Figures

Figure 1

Figure 1
<p>The EDX for elemental composition of the selected CHC.</p>
Full article ">Figure 2
<p>SEM for the selected CHC.</p>
Full article ">Figure 3
<p>The FTIR spectrum for the raw composite selected residue.</p>
Full article ">Figure 4
<p>The FTIR spectrum of the composite calcined heterogeneous catalysts.</p>
Full article ">Figure 5
<p>XRD spectrum for CHC.</p>
Full article ">
22 pages, 4049 KiB  
Article
Synthesis of a Novel Magnetic Biochar from Lemon Peels via Impregnation-Pyrolysis for the Removal of Methyl Orange from Wastewater
by Samah Daffalla, Enshirah Da’na, Amel Taha and Mohamed R. El-Aassar
Magnetochemistry 2024, 10(12), 95; https://doi.org/10.3390/magnetochemistry10120095 - 29 Nov 2024
Viewed by 323
Abstract
This research examined the elimination of methyl orange (MO) utilizing a novel magnetic biochar adsorbent (MLPB) derived from lemon peels via an impregnation-pyrolysis method. Material characterization was conducted using SEM, XRD, TGA, FTIR, and nitrogen adsorption isotherms. SEM-EDX analysis indicates that MLPB is [...] Read more.
This research examined the elimination of methyl orange (MO) utilizing a novel magnetic biochar adsorbent (MLPB) derived from lemon peels via an impregnation-pyrolysis method. Material characterization was conducted using SEM, XRD, TGA, FTIR, and nitrogen adsorption isotherms. SEM-EDX analysis indicates that MLPB is a homogeneous and porous composite comprising Fe, O, and C, with iron oxide uniformly dispersed throughout the material. Also, MLPB is porous with an average pore diameter of 4.65 nm and surface area value (111.45 m2/g). This study evaluated pH, MO concentration, and contact time to analyze the adsorption process, kinetics, and isothermal behavior. Under optimal conditions, MLPB was able to remove MO dye from aqueous solutions with an efficiency of 90.87%. Results showed optimal MO removal at pH 4, suggesting a favorable electrostatic interaction between the adsorbent and dye. To ascertain the adsorption kinetics, the experimental findings were compared using several adsorption models, first- and second-orders, and intra-particle diffusion. According to the findings, the pseudo-second-order model described the adsorption kinetic promoting the formation of the chemisorption phase well. Modeling of intra-particle diffusion revealed that intra-particle diffusion is not the only rate-limiting step. A study involving isothermal systems showed that Langmuir is a good representation of experimental results; the maximum adsorption capacity of MLPB was 17.21 mg/g. According to the results, after four cycles of regeneration, the produced magnetic material regained more than 88% of its adsorption ability. Full article
(This article belongs to the Special Issue Applications of Magnetic Materials in Water Treatment)
Show Figures

Figure 1

Figure 1
<p>MLPB and LP’s FT-IR spectrum.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) LP, (<b>b</b>) MLPB, and (<b>c</b>–<b>e</b>) C, O, and Fe; elemental distribution over the same region mapped per each element.</p>
Full article ">Figure 3
<p>XRD patterns of MLPB.</p>
Full article ">Figure 4
<p>TGA curves of LP and MLPB.</p>
Full article ">Figure 5
<p>The MLPB sample’s (<b>a</b>) N<sub>2</sub> absorption/desorption isotherms and (<b>b</b>) pore size distribution plot.</p>
Full article ">Figure 6
<p>Impact of initial pH solution on MO’s ability to adsorb onto MLPB (20 mg/L starting Mo concentration, 0.1 g adsorbent dosage).</p>
Full article ">Figure 7
<p>The impact of the MLPB dosage on the MO adsorption capacity at pH 3 and the starting Mo concentration of 20 mg/L.</p>
Full article ">Figure 8
<p>Effect of initial dye concentration of adsorption of MO dye unto MPLB at pH 3, 25 °C, and 0.1 g adsorbent dose.</p>
Full article ">Figure 9
<p>Impact of initial dye concentration and contact duration on MO dye adsorption onto MPLB at pH 3, 25 °C, and 0.1 g adsorbent dosage.</p>
Full article ">Figure 10
<p>(<b>a</b>) Pseudo-first-order, (<b>b</b>) pseudo-second-order, and (<b>c</b>) Intraparticle diffusion kinetics plots at different initial MO concentrations.</p>
Full article ">Figure 11
<p>(<b>a</b>) Langmuir, (<b>b</b>) for Freundlich, and (<b>c</b>) Dubinin–Radushkevich (D–R) isotherms plots.</p>
Full article ">Figure 12
<p>Recycling of MLPB for multiple adsorption-desorption of MO.</p>
Full article ">Figure 13
<p>Breakthrough for MO at initial concentrations of 20, 50, and 80 mg/L.</p>
Full article ">Figure 14
<p>MLPB’s FTIR before and after MO adsorption.</p>
Full article ">Figure 15
<p>Suggested mechanism for adsorption of MO on MLPB.</p>
Full article ">
21 pages, 8148 KiB  
Article
Green Synthesis of Titanium Dioxide Nanoparticles Using Maerua oblongifolia Root Bark Extract: Photocatalytic Degradation and Antibacterial Activities
by Mamo Dikamu Dilika, Gada Muleta Fanta and Tomasz Tański
Materials 2024, 17(23), 5835; https://doi.org/10.3390/ma17235835 - 28 Nov 2024
Viewed by 528
Abstract
The root bark extract of the Maerua oblongifolia plant in the green synthesis of titanium dioxide nanoparticles (TiO2 NPs) for photocatalytic degradation of toxic pollutants and antibacterial activities was implemented in this study. The root bark extract served as a novel capping [...] Read more.
The root bark extract of the Maerua oblongifolia plant in the green synthesis of titanium dioxide nanoparticles (TiO2 NPs) for photocatalytic degradation of toxic pollutants and antibacterial activities was implemented in this study. The root bark extract served as a novel capping and reducing agent for the first time. Characterization of the TiO2 NPs was conducted by using visual observation, ultraviolet visible spectrometry (UV-Vis), Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), and scanning electron microscopy (SEM) techniques, confirming their successful synthesis. The TiO2 NPs exhibited maximum absorbance at 323 nm and an average particle size of 19.58 nm; the conjugations and existences of Ti-O and OH vibrational bands were revealed by the FTIR spectrum. The photocatalytic activities of the TiO2 NPs were investigated by using solar irradiation as an energy source for aqueous solutions of methyl orange (MO) and methylene blue (MB) dyes. The TiO2 NPs showed strong photocatalytic activities by degrading 97.23% MB and 91.8% MO under optimized conditions. Degradation behavior was investigated by isotherms and kinetics models, with the Langmuir isotherms (R2: 0.996, 0.979) and Langmuir–Hinshelwood (R2: 0.998, 0.997) highest correlation coefficients for MB and MO, respectively. Moreover, the antibacterial efficacy of the green-synthesized TiO2 NPs and the results indicated higher antibacterial activities on Gram-negative bacteria (27 ± 0.52). Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the work flow of this study.</p>
Full article ">Figure 2
<p>(<b>a</b>). Synthesized TiO<sub>2</sub> NPs; UV-Vis spectrum (<b>b</b>). XRD pattern of the TiO<sub>2</sub> NPs synthesized using saponin-containing extract (<b>c</b>). XRD pattern of the TiO<sub>2</sub> NPs synthesized without saponin extract. (<b>d</b>) FTIR spectra of both the <span class="html-italic">M. oblongifolia</span> root bark extract and green-synthesized TiO<sub>2</sub> NPs.</p>
Full article ">Figure 3
<p>Possible reaction mechanism for the formation of TiO<sub>2</sub> NPs in the presence of the hydroxyl group (-OH) of the leaf extract of <span class="html-italic">Jatropha curcas</span> L. as a capping agent [<a href="#B33-materials-17-05835" class="html-bibr">33</a>].</p>
Full article ">Figure 4
<p>Mechanism of the bioreduction of titanyl hydroxide to TiO<sub>2</sub> NPs using <span class="html-italic">Euphorbia hetarade Jaub</span> root extract [<a href="#B34-materials-17-05835" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>Green-synthesized TiO<sub>2</sub> NPs’ bandgap energy.</p>
Full article ">Figure 6
<p>SEM results of the TiO<sub>2</sub> NPs. (<b>a</b>,<b>b</b>) indicate before and after isolations of saponins and (<b>c</b>,<b>d</b>) indicate the particulate size distributions with and without saponin extract, respectively.</p>
Full article ">Figure 7
<p>Photocatalytic degradation mechanism over TiO<sub>2</sub> NPs under sunlight irradiation.</p>
Full article ">Figure 8
<p>Kinetic study of both MB and MO at optimum conditions.</p>
Full article ">Figure 9
<p>Three-dimensional plot and counterplot for the effects of irradiation time and catalyst dose.</p>
Full article ">Figure 10
<p>Three-dimensional plot and counterplot for the influence of catalytic dose and pH of the solutions.</p>
Full article ">Figure 11
<p>Three-dimensional plot and counterplot for effects of catalytic dose and initial dye concentration.</p>
Full article ">Figure 12
<p>Langmuir–Hinshelwood kinetic degradation models for MB and MO dyes.</p>
Full article ">Figure 13
<p>Kinetic models for the degradations of MB and MO.</p>
Full article ">Figure 14
<p>Reusability runs of the catalyst for both MB and MO dyes.</p>
Full article ">Figure 15
<p>Antibacterial activity (zone of inhibition): images of TiO<sub>2</sub> nanoparticles, the positive control, and the plant extract against the pathogens <span class="html-italic">S. aureus</span> and <span class="html-italic">E. coli</span> at different concentrations.</p>
Full article ">
10 pages, 2631 KiB  
Article
Improvement of Monacolin K and Pigment Production in Monascus by 5-Azacytidine
by Chan Zhang, Haijiao Wang, Qing Sun, Arzugul Ablimit, Huijun Dong, Congcong Wang, Duchen Zhai, Bobo Zhang, Wenlin Hu, Chengjian Liu and Chengtao Wang
J. Fungi 2024, 10(12), 819; https://doi.org/10.3390/jof10120819 - 26 Nov 2024
Viewed by 428
Abstract
Monascus species are known to produce various secondary metabolites with polyketide structures, including Monacolins, pigments, and citrinin. This study investigates the effects of 5-azacytidine on Monascus M1 and RP2. The dry weight, red, yellow, and orange pigment values, and Monacolin K yield of [...] Read more.
Monascus species are known to produce various secondary metabolites with polyketide structures, including Monacolins, pigments, and citrinin. This study investigates the effects of 5-azacytidine on Monascus M1 and RP2. The dry weight, red, yellow, and orange pigment values, and Monacolin K yield of both Monascus strains were measured, and their hyphae observed through electron microscopy. The experimental group showed higher dry weights and pigment values than the control group for both strains. However, Monacolin K production increased substantially only for Monascus M1. Electron micrographs revealed surface wrinkles and large protrusions in both strains after 5-azacytidine treatment. As a potent DNA methylation-promoting agent, 5-azacytidine is very useful for epigenetic and cancer biology studies and for studying secondary metabolism in fungi. Full article
(This article belongs to the Special Issue Monascus spp. and Their Relative Products)
Show Figures

Figure 1

Figure 1
<p>Effects of 5-azacytidine on <span class="html-italic">Monascus</span> M1 and <span class="html-italic">Monascus</span> RP2. (<b>A</b>) Yield of Monacolin K of <span class="html-italic">Monascus</span> M1 at different addition times. (<b>B</b>) Red pigment value of <span class="html-italic">Monascus</span> RP2 at different addition times. (<b>C</b>) Yield of Monacolin K of <span class="html-italic">Monascus</span> M1 at different concentrations of 5-azacytidine. (<b>D</b>) Red pigment value of <span class="html-italic">Monascus</span> RP2 at different concentrations of 5-azacytidine, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 2
<p>Effects of 5-azacytidine on <span class="html-italic">Monascus</span> M1 and <span class="html-italic">Monascus</span> RP2. (<b>A</b>) Dry weight of <span class="html-italic">Monascus</span> M1. (<b>B</b>) Dry weight of <span class="html-italic">Monascus</span> RP2, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 3
<p>Effects of 5-azacytidine on the pigment value in <span class="html-italic">Monascus</span> M1 and <span class="html-italic">Monascus</span> RP2. (<b>A</b>) The red pigment value of <span class="html-italic">Monascus</span> M1, (<b>B</b>) the red pigment value of <span class="html-italic">Monascus</span> RP2, (<b>C</b>) the orange pigment value of <span class="html-italic">Monascus</span> M1, (<b>D</b>) the orange pigment value of <span class="html-italic">Monascus</span> RP2, (<b>E</b>) the yellow pigment value of <span class="html-italic">Monascus</span> M1, and (<b>F</b>) the yellow pigment value of <span class="html-italic">Monascus</span> RP2, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 4
<p>Yield of Monacolin K of <span class="html-italic">Monascus</span> M1, * <span class="html-italic">p</span> &lt; 0.05.</p>
Full article ">Figure 5
<p>Scanning electron micrographs showing the morphology of <span class="html-italic">Monascus</span> RP2 (<b>A</b>–<b>D</b>) and <span class="html-italic">Monascus</span> M1 (<b>E</b>–<b>H</b>) at 8 days in different culture media with different magnification factors (5000× and 10,000×, respectively). <span class="html-italic">Monascus</span> RP2 (<b>A</b>,<b>C</b>) in the original medium, <span class="html-italic">Monascus</span> M1 (<b>E</b>,<b>G</b>) in the original medium, <span class="html-italic">Monascus</span> RP2 (<b>B</b>,<b>D</b>) in the medium containing 5-azacytidine, and <span class="html-italic">Monascus</span> M1 (<b>F</b>,<b>H</b>) in the medium containing 5-azacytidine. Arrows indicate wrinkles.</p>
Full article ">
16 pages, 2449 KiB  
Article
Identification of Cherry Tomato Volatiles Using Different Electron Ionization Energy Levels
by Dalma Radványi, László Csambalik, Dorina Szakál and Attila Gere
Molecules 2024, 29(23), 5567; https://doi.org/10.3390/molecules29235567 - 25 Nov 2024
Viewed by 407
Abstract
A comprehensive analysis of the volatile components of 11 different cherry tomato pastes (Tesco Extra, Orange, Zebra, Yellow, Round Netherland, Mini San Marzano, Spar truss, Tesco Sunstream, Paprikakertész, Mc Dreamy, and Tesco Eat Fresh) commercially available in Hungary was performed. In order to [...] Read more.
A comprehensive analysis of the volatile components of 11 different cherry tomato pastes (Tesco Extra, Orange, Zebra, Yellow, Round Netherland, Mini San Marzano, Spar truss, Tesco Sunstream, Paprikakertész, Mc Dreamy, and Tesco Eat Fresh) commercially available in Hungary was performed. In order to ensure the reliability and accuracy of the measurement, the optimal measurement conditions were first determined. SPME (solid-phase microextraction) fiber coating, cherry tomato paste treatment, and SPME sampling time and temperature were optimized. CAR/PDMS (carboxen/polydimethylsiloxane) fiber coating with a film thickness of 85 µm is suggested at a 60 °C sampling temperature and 30 min extraction time. A total of 64 common compounds was found in the prepared, mashed cherry tomato samples, in which 59 compounds were successfully identified. Besides the already published compounds, new, cherry tomato-related compounds were found, such as 3 methyl 2 butenal, heptenal, Z-4-heptenal, E-2-heptenal, E-carveol, verbenol, limonene oxide, 2-decen-1-ol, Z-4-decen-1-al, caryophyllene oxide, and E,E-2,4-dodecadienal. Supervised and unsupervised classification methods have been used to classify the tomato varieties based on their volatiles, which identified 16 key components that enable the discrimination of the samples with a high accuracy. Full article
(This article belongs to the Special Issue Extraction and Analysis of Natural Products in Food—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Comparison of SPME fiber coatings in the case of captured tomato volatiles. (<b>b</b>) Examination of the effect of seven different extraction temperatures (10, 25, 30, 35, 50, 60, and 80 min) on captured VOCs from tomatoes. (<b>c</b>) Examination of the effect of six different extraction times (10, 20, 30, 40, 60, and 120 min) on captured VOCs from tomatoes. (<b>d</b>) Evaluation of the effect of solvent addition (methanol and water).</p>
Full article ">Figure 2
<p>Mass spectra of hexanal at 70 eV, 60 eV, 50 eV, 40 eV, 20 eV, 10 eV, and 5 eV.</p>
Full article ">Figure 3
<p>Mass spectra of 2H-1b,4-Ethanopentaleno[1,2-b]oxirene,hexahydro-(1aa,1bb,4b,4aa,5aa) at 70 eV and 5 eV. Match factor: 92.8%, structure: C<sub>10</sub>H<sub>14</sub>O, molecular weight: 150.1, and CAS number: 117221-80-4. TIC: total ion chromatogram.</p>
Full article ">Figure 4
<p>Agglomerative hierarchical clustering using the Euclidean distance and Ward’s method. Dashed line represents an arbitrary cut of the dendrogram. Colors and numbers represent the mashed tomato samples (1: Tesco Extra, 2: Orange, 3: Zebra, 4: Yellow, 5: Round Netherland, 6: Mini San Marzano, 7: Spar truss, 8: Tesco Sunstream, 9: Paprikakertész, 10: Mc Dreamy, and a11: Tesco Eat Fresh), while letters represent the repetitions (a, b, and c).</p>
Full article ">Figure 5
<p>The discriminant analysis runs on principal component analysis loadings. The first discriminant function accounts for 97.76% of variance. Colors and numbers represent the mashed tomato samples (1: Tesco Extra, 2: Orange, 3: Zebra, 4: Yellow, 5: Round Netherland, 6: Mini San Marzano, 7: Spar truss, 8: Tesco Sunstream, 9: Paprikakertész, 10: Mc Dreamy, and 11: Tesco Eat Fresh), while letters represent the repetitions (a, b, and c). D: discriminant function.</p>
Full article ">
9 pages, 703 KiB  
Article
Production of Pectic Oligosaccharides from Citrus Peel via Steam Explosion
by Toni-Ann Martorano, Kyle L. Ferguson, Randall G. Cameron, Wei Zhao, Arland T. Hotchkiss, Hoa K. Chau and Christina Dorado
Foods 2024, 13(23), 3738; https://doi.org/10.3390/foods13233738 - 22 Nov 2024
Viewed by 747
Abstract
Steam explosion (STEX) of peel from commercially juice-extracted oranges was used to convert peel pectin into pectic oligosaccharides (POSs). Surprisingly uniform populations, based on the polydispersity index (PDI; weight-average molecular weight (Mw)/number-average molecular weight (Mn)) of POSs, were obtained [...] Read more.
Steam explosion (STEX) of peel from commercially juice-extracted oranges was used to convert peel pectin into pectic oligosaccharides (POSs). Surprisingly uniform populations, based on the polydispersity index (PDI; weight-average molecular weight (Mw)/number-average molecular weight (Mn)) of POSs, were obtained from the Hamlin and Valencia varieties of Citrus sinensis. The POSs from Hamlin and Valencia peel had PDI values of (1.23 ± 0.01, 1.24 ± 0.1), respectively. The Mw values for these samples were 14.9 ± 0.2 kDa for Hamlin, and 14.5 ± 0.1 kDa for Valencia, respectively. The degree of methyl-esterification (DM) was 69.64 ± 3.18 for Hamlin and 65.51 ± 1.61 for Valencia. The composition of the recovered POSs was dominated by galacturonic acid, ranging from 89.1% to 99.6% of the major pectic sugars. Only the Hamlin sample had a meaningful amount of rhamnose present, indicating the presence of an RG I domain. Even so, the Hamlin sample’s degree of branching (DBr) was very low (2.95). Full article
Show Figures

Figure 1

Figure 1
<p>High-performance size-exclusion chromatography analysis of the Hamlin (DM-D4) (<b>A</b>) and Valencia (DM-D6) (<b>B</b>) varieties; superimposed calibration curve of Hamlin and Valencia (DM-D4, DM-D6) (<b>C</b>). In <a href="#foods-13-03738-f001" class="html-fig">Figure 1</a>A,B, HPSEC detectors were light scattering at 90 °C (<span style="color:red">-</span>), differential pressure viscometer (-), refractive index (<span style="color:#0070C0">-</span>), and ultraviolet absorption at 280 nm (<span style="color:#92D050">-</span>). In <a href="#foods-13-03738-f001" class="html-fig">Figure 1</a>C, HPSEC detectors were light scattering at 90 °C, (DM-D4) and HPSEC detectors were light scattering at 90 °C, DM-D6.</p>
Full article ">
16 pages, 7984 KiB  
Article
Efficient Catalytic Reduction of Organic Pollutants Using Nanostructured CuO/TiO2 Catalysts: Synthesis, Characterization, and Reusability
by Mariyem Abouri, Abdellah Benzaouak, Fatima Zaaboul, Aicha Sifou, Mohammed Dahhou, Mohammed Alaoui El Belghiti, Khalil Azzaoui, Belkheir Hammouti, Larbi Rhazi, Rachid Sabbahi, Mohammed M. Alanazi and Adnane El Hamidi
Inorganics 2024, 12(11), 297; https://doi.org/10.3390/inorganics12110297 - 19 Nov 2024
Viewed by 751
Abstract
The catalytic reduction of organic pollutants in water is a critical environmental challenge due to the persistent and hazardous nature of compounds like azo dyes and nitrophenols. In this study, we synthesized nanostructured CuO/TiO2 catalysts via a combustion technique, followed by calcination [...] Read more.
The catalytic reduction of organic pollutants in water is a critical environmental challenge due to the persistent and hazardous nature of compounds like azo dyes and nitrophenols. In this study, we synthesized nanostructured CuO/TiO2 catalysts via a combustion technique, followed by calcination at 700 °C to achieve a rutile-phase TiO2 structure with varying copper loadings (5–40 wt.%). The catalysts were characterized using X-ray diffraction (XRD), attenuated total reflectance-Fourier transform infrared (ATR–FTIR) spectroscopy, thermogravimetric analysis-differential thermal analysis (TGA–DTA), UV-visible diffuse reflectance spectroscopy (DRS), and scanning electron microscopy with energy-dispersive X-ray spectroscopy (SEM–EDS). The XRD results confirmed the presence of the crystalline rutile phase in the CuO/TiO2 catalysts, with additional peaks indicating successful copper oxide loading onto TiO2. The FTIR spectra confirmed the presence of all the functional groups in the prepared samples. SEM images revealed irregularly shaped copper oxide and agglomerated TiO2 particles. The DRS results revealed improved optical properties and a decreased bandgap with increased Cu content, and 4-Nitrophenol (4-NP) and methyl orange (MO), which were chosen for their carcinogenic, mutagenic, and nonbiodegradable properties, were used as model organic pollutants. Catalytic activities were tested by reducing 4-NP and MO with sodium borohydride (NaBH4) in the presence of a CuO/TiO2 catalyst. Following the in situ reduction of CuO/TiO2, Cu (NPs)/TiO2 was formed, achieving 98% reduction of 4-NP in 480 s and 98% reduction of MO in 420 s. The effects of the NaBH4 concentration and catalyst mass were investigated. The catalysts exhibited high stability over 10 reuse cycles, maintaining over 96% efficiency for MO and 94% efficiency for 4-NP. These findings demonstrate the potential of nanostructured CuO/TiO2 catalysts for environmental remediation through efficient catalytic reduction of organic pollutants. Full article
(This article belongs to the Special Issue New Advances into Nanostructured Oxides, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of different wt.% CuO/TiO<sub>2</sub> samples.</p>
Full article ">Figure 2
<p>XRD patterns of 40 wt.% CuO/TiO<sub>2</sub> samples calcined at different temperatures.</p>
Full article ">Figure 3
<p>XRD patterns of 40 wt.% CuO/TiO<sub>2</sub> before and after the reduction experiment.</p>
Full article ">Figure 4
<p>Thermal curves of the as-prepared 40 wt.% CuO/TiO<sub>2</sub> without calcination under air flow at 10 °C min<sup>−1</sup>.</p>
Full article ">Figure 5
<p>Infrared spectra of the wt.% CuO/TiO<sub>2</sub> samples.</p>
Full article ">Figure 6
<p>(<b>a</b>) UV–vis diffuse reflectance spectra of TiO<sub>2</sub> and wt.% CuO/TiO<sub>2</sub> and (<b>b</b>) Tauc plot.</p>
Full article ">Figure 7
<p>SEM micrographs of 40 wt.% CuO/TiO<sub>2</sub>.</p>
Full article ">Figure 8
<p>EDS Spectrum of 40 wt.% CuO/TiO<sub>2</sub>.</p>
Full article ">Figure 9
<p>EDX mapping of the 40 wt.% CuO/TiO<sub>2</sub> catalyst.</p>
Full article ">Figure 10
<p>Reduction performance of the wt.% CuO/TiO<sub>2</sub> catalysts for 4-NP (<b>a</b>) and MO (<b>b</b>).</p>
Full article ">Figure 11
<p>UV–vis spectra of the reduction of (<b>a</b>) 4-NP and (<b>b</b>) MO by NaBH<sub>4</sub> without the catalyst and absorbance spectra of (<b>c</b>) 4-NP and (<b>d</b>) MO in the presence of NaBH<sub>4</sub> and CuO/TiO<sub>2</sub>.</p>
Full article ">Figure 11 Cont.
<p>UV–vis spectra of the reduction of (<b>a</b>) 4-NP and (<b>b</b>) MO by NaBH<sub>4</sub> without the catalyst and absorbance spectra of (<b>c</b>) 4-NP and (<b>d</b>) MO in the presence of NaBH<sub>4</sub> and CuO/TiO<sub>2</sub>.</p>
Full article ">Figure 12
<p>First-order plot for the catalytic reduction of (<b>a</b>) 4-NP and (<b>b</b>) MO. Experimental conditions: 40 wt.% CuO/TiO<sub>2</sub> dose = 0.04 g/L, [NaBH<sub>4</sub>] = 14 mM, and [4-NP or MO] = 0.35 mM.</p>
Full article ">Figure 13
<p>Schematic reactions of (<b>a</b>) 4-NP and (<b>b</b>) MO, by NaBH<sub>4</sub> in the presence of Cu(NPs)/TiO<sub>2</sub>.</p>
Full article ">Figure 14
<p>Pollutant removal versus time plots for catalytic reduction using 40 wt.% CuO/TiO<sub>2</sub>, showing the effect of catalyst dose on the reduction of (<b>a</b>) 4-NP and (<b>b</b>) MO, and the effect of NaBH<sub>4</sub> concentration on the reduction of (<b>c</b>) 4-NP and (<b>d</b>) MO.</p>
Full article ">Figure 15
<p>Recycling of CuO/TiO<sub>2</sub> in the hydrogenation of (<b>a</b>) 4-NP and (<b>b</b>) MO.</p>
Full article ">Figure 16
<p>Synthesis procedure for the catalysts.</p>
Full article ">Figure 17
<p>Catalytic reduction experiments of organic pollutants.</p>
Full article ">
18 pages, 1783 KiB  
Article
New Polyhydroxysteroid Glycosides with Antioxidant Activity from the Far Eastern Sea Star Ceramaster patagonicus
by Timofey V. Malyarenko, Viktor M. Zakharenko, Alla A. Kicha, Arina I. Ponomarenko, Igor V. Manzhulo, Anatoly I. Kalinovsky, Roman S. Popov, Pavel S. Dmitrenok and Natalia V. Ivanchina
Mar. Drugs 2024, 22(11), 508; https://doi.org/10.3390/md22110508 - 10 Nov 2024
Viewed by 1066
Abstract
Four new glycosides of polyhydroxysteroids, ceramasterosides A, B, D, and E (14), and two previously known compounds, ceramasteroside C1 (5) and attenuatoside B-I (6), were isolated from an extract of a deep-sea sea star [...] Read more.
Four new glycosides of polyhydroxysteroids, ceramasterosides A, B, D, and E (14), and two previously known compounds, ceramasteroside C1 (5) and attenuatoside B-I (6), were isolated from an extract of a deep-sea sea star species, the orange cookie star Ceramaster patagonicus. The structures of 14 were elucidated by the extensive NMR and ESIMS methods. Steroid monoglycosides 1 and 2 had a common 3β,6α,8,15β,16β-pentahydroxysteroid nucleus and a C–29 oxidized stigmastane side chain and differed from each other only in monosaccharide residues. Ceramasteroside A (1) contained 3-O-methyl-4-O-sulfated β-D-xylopyranose, while ceramasteroside B (2) had 3-O-methyl-4-O-sulfated β-D-glucopyranose, recorded from starfish-derived steroid glycosides for the first time. Their biological activity was studied using a model of lipopolysaccharide-induced (LPS) inflammation in a SIM-A9 murine microglial cell line. During the LPS-induced activation of microglial cells, 1, 3, and 5, at a non-toxic concentration of 1 µM, showed the highest efficiency in reducing the production of intracellular NO, while 4 proved to be most efficient in reducing the extracellular nitrite production. All the test compounds reduced the LPS-induced malondialdehyde (MDA) production. The in vitro experiments have demonstrated, for the first time, the antioxidant activity of the compounds under study. Full article
(This article belongs to the Special Issue Biologically Active Compounds from Marine Invertebrates 2025)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The structures of <b>1</b>−<b>6</b> isolated from <span class="html-italic">C. patagonicus</span>.</p>
Full article ">Figure 2
<p><sup>1</sup>H-<sup>1</sup>H COSY and key HMBC correlations of <b>1</b>–<b>4</b>.</p>
Full article ">Figure 3
<p>Key ROESY correlations of <b>1</b>–<b>4</b>.</p>
Full article ">Figure 4
<p>(<b>A</b>) Images of microglial cells exposed to lipopolysaccharide (LPS) and <b>1</b>–<b>6</b> at a concentration of 10 µM. (<b>B</b>) The cytotoxic effect of <b>1</b>–<b>6</b> on mouse microglial cell line SIM-A9 determined by MTS assay. Data are mean ± SEM; <span class="html-italic">n</span> = 8 (number of samples analyzed); * <span class="html-italic">p</span> &lt; 0.05 vs. control (CTL) (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 5
<p>(<b>A</b>) Morphological changes in microglia under the exposure to LPS and <b>1</b>–<b>6</b> at a concentration of 10 µM. Antioxidant activity of <b>1</b>–<b>6</b> against LPS-induced production of NO (<b>B</b>) and nitrites (<b>C</b>) in mouse microglial cell line SIM-A9. Data are mean ± SEM; <span class="html-italic">n</span> = 8 (number of samples analyzed)’ * <span class="html-italic">p</span> &lt; 0.05 vs. CTL, <sup>+</sup> <span class="html-italic">p</span> &lt; 0.05 vs. LPS (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">Figure 6
<p>Analysis of MDA production in SIM-A9 cell line under the exposure to LPS and <b>1</b>–<b>6</b> at a concentration of 1 µM. Data are mean ± SEM; <span class="html-italic">n</span> = 8 (number of samples analyzed); * <span class="html-italic">p</span> &lt; 0.05 vs. CTL; <sup>+</sup> <span class="html-italic">p</span> &lt; 0.05 vs. LPS (Student’s <span class="html-italic">t</span>-test).</p>
Full article ">
16 pages, 4716 KiB  
Article
Photocatalytic Degradation of Four Organic Dyes Present in Water Using ZnO Nanoparticles Synthesized with Green Synthesis Using Ambrosia ambrosioides Leaf and Root Extract
by Martin Medina-Acosta, Manuel J. Chinchillas-Chinchillas, Horacio E. Garrafa-Gálvez, Caree A. Garcia-Maro, Carlos A. Rosas-Casarez, Eder Lugo-Medina, Priscy A. Luque-Morales and Carlos A. Soto-Robles
Processes 2024, 12(11), 2456; https://doi.org/10.3390/pr12112456 - 6 Nov 2024
Viewed by 904
Abstract
Currently, several organic dyes found in wastewater cause severe contamination problems for flora, fauna, and people in direct contact with them. This research proposes an alternative for the degradation of polluting dyes using ZnO nanoparticles (NPs) synthesized by an ecological route using leaf [...] Read more.
Currently, several organic dyes found in wastewater cause severe contamination problems for flora, fauna, and people in direct contact with them. This research proposes an alternative for the degradation of polluting dyes using ZnO nanoparticles (NPs) synthesized by an ecological route using leaf and root extracts of Ambrosia ambrosioides as a reducing agent (with a weight/volume ratio = 4%). Scanning Electron Microscopy (SEM) was used to determine the morphology, showing an agglomeration of cluster-shaped NPs. Using Transmission Electron Microscopy (TEM), different sizes of NPs ranging from 5 to 56 nm were observed for both synthesized NPs. The composition and structure of the nanomaterial were analyzed by infrared spectroscopy (FT-IR) and X-ray diffraction (XRD), showing as a result that the NPs have a wurtzite-like crystalline structure with crystallite sizes around 32–37 nm for both samples. Additionally, the bandgap of the NPs was calculated using Ultraviolet Visible Spectroscopy (UV–Vis), determining values of 2.82 and 2.70 eV for the NPs synthesized with leaf and root, respectively. Finally, thermogravimetric analysis demonstrated that the nanoparticles contained an organic part after the green synthesis process, with high thermal stability for both samples. Photocatalytic analysis showed that these nanomaterials can degrade four dyes under UV irradiation, reaching 90% degradation for methylene blue (MB), methyl orange (MO) and Congo red (CR) at 60, 100 and 60 min, respectively, while for methyl red (MR) almost 90% degradation was achieved at 140 min of UV irradiation. These results demonstrate that it is effective to use Ambrosia ambrosioides root and leaf extracts as a reducing agent for the formation of ZnO NPs, also evidencing their favorable application in the photocatalytic degradation of these four organic dyes. Full article
Show Figures

Figure 1

Figure 1
<p>FT-IR spectra of nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides</span>.</p>
Full article ">Figure 2
<p>XRD spectra of ZnO nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides</span> using: (<b>a</b>) root and (<b>b</b>) sheet.</p>
Full article ">Figure 3
<p>UV–Vis analysis of ZnO nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides</span>: (<b>a</b>,<b>c</b>) absorbance of ZnO nanoparticles synthesized with root and leaf, respectively, and (<b>b</b>,<b>d</b>) bandgap calculation of ZnO nanoparticles synthesized with root and leaf, respectively.</p>
Full article ">Figure 4
<p>SEM morphology study of ZnO nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides:</span> (<b>a</b>,<b>b</b>) Morphology of AA_R_ZnO nanoparticles, (<b>c</b>) size distribution of AA_R_ZnO, (<b>d</b>,<b>e</b>) Morphology of AA_S_ZnO nanoparticles and (<b>f</b>) size distribution of AA_S_ZnO.</p>
Full article ">Figure 5
<p>TEM morphology study of ZnO nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides</span> extracts. (<b>a</b>,<b>b</b>) Morphology of AA_R_ZnO nanoparticles, (<b>c</b>) SAED of AA_R_ZnO, (<b>d</b>,<b>e</b>) Morphology of AA_S_ZnO nanoparticles and (<b>f</b>) SAED of AA_S_ZnO.</p>
Full article ">Figure 6
<p>TGA/DSC of the nanoparticles synthesized using <span class="html-italic">Ambrosia ambrosioides</span> extracts.</p>
Full article ">Figure 7
<p>Photocatalytic degradations of (<b>a</b>) MB, (<b>b</b>) MO, (<b>c</b>) CR and (<b>d</b>) MR using ZnO nanoparticles synthesized with <span class="html-italic">Ambrosia ambrosioides</span> extracts.</p>
Full article ">Figure 8
<p>Mechanism of photocatalytic degradation proposed by this research.</p>
Full article ">Figure 9
<p>Green synthesis process of ZnO nanoparticles using <span class="html-italic">Ambrosia ambrosioides</span>.</p>
Full article ">
19 pages, 4587 KiB  
Article
Sustainable Activated Carbon from Agricultural Waste: A Study on Adsorption Efficiency for Humic Acid and Methyl Orange Dyes
by Zahia Tigrine, Ouassila Benhabiles, Leila Merabti, Nadia Chekir, Mounir Mellal, Salaheddine Aoudj, Nora Amele Abdeslam, Djilali Tassalit, Seif El Islam Lebouachera and Nadjib Drouiche
Sustainability 2024, 16(21), 9308; https://doi.org/10.3390/su16219308 - 26 Oct 2024
Viewed by 1137
Abstract
In this study, porous activated carbon was produced from coffee waste and used as an effective adsorbent for the removal of humic acid (HA) from seawater and methyl orange (MO) dye from aqueous solutions. Phosphoric acid H3PO4 was used as [...] Read more.
In this study, porous activated carbon was produced from coffee waste and used as an effective adsorbent for the removal of humic acid (HA) from seawater and methyl orange (MO) dye from aqueous solutions. Phosphoric acid H3PO4 was used as an activating agent for the chemical activation of these agricultural wastes. The characterization of the activated carbon obtained using a scanning electron microscope (SEM), Fourier-transform infrared (FTIR) spectroscopy analysis, X-ray diffraction (XRD) and the Brunauer–Emmett–Teller (BET) method revealed that the activated carbon products exhibited high porosity and the formation of various functional groups. The effects of different parameters were examined using batch adsorption experiments, such as the adsorbent masses, pH, initial pollutant concentration and contact time. The results show that the performance increased with an increased adsorbent mass (up to 0.25 g/L) and decreased initial concentration of the adsorbent tested. On the other hand, this study clearly showed that the adsorption efficiency of the MO on the raw spent coffee grounds (SCGs) waste was around 43%, while no removal was observed for the humic acid. The experiments demonstrated that the activated carbon synthesized from the used coffee grounds (the efficiency was compared with commercial activated carbon (CAC) with a difference of 13%) was a promising alternative to commercially available adsorbents for the removal of humic acid from seawater. To understand and elucidate the adsorption mechanism, various isothermal and kinetic models were studied. The adsorption capacity was analyzed by fitting experimental data to these models. The experimental data for methyl orange dyes were analyzed using Langmuir and Freundlich isothermal models. The Freundlich isotherm model provided a superior fit to the equilibrium data, as indicated by a higher correlation coefficient (R2) than that of the Langmuir model. The maximum adsorption was observed at pH 3. The Freundlich adsorption capacity was found to be 333 mg/g adsorbent. The PAC showed a high adsorption capacity for the MO and HA. The PAC showed the highest adsorption capacities for the HA and MO compared with the other adsorbents used (SCGs and CAC) and would be a good material to increase the adsorption efficiency for humic acid removal in the seawater pretreatment process. In addition, the prepared AC BET surface area was 520.40 m2/g, suggesting a high adsorption capacity. This makes the material potentially suitable for various applications that require a high surface area. These results indicate that high-quality sustainable activated carbon can be efficiently produced from coffee waste, making it suitable for a wide range of adsorbent applications targeting various pollutants. Full article
Show Figures

Figure 1

Figure 1
<p>Chemical structures, (<b>a</b>) methyl orange (MO) and(<b>b</b>) humic acid (HA).</p>
Full article ">Figure 2
<p>SEM images of materials taken at different magnifications before and after chemical activation with H<sub>3</sub>PO<sub>4</sub> and pyrolysis at 550 °C, (<b>a</b>) Spent coffee grounds (SCGs) before chemical activation and (<b>b</b>) Prepared activated carbon from SCGs (after chemical activation).</p>
Full article ">Figure 3
<p>Fourier-transform infrared (FTIR) spectra of materials: (<b>a</b>) raw spent coffee grounds and (<b>b</b>) prepared activated carbon.</p>
Full article ">Figure 4
<p>X-ray diffraction (XRD) of materials used: SCGs and PAC.</p>
Full article ">Figure 5
<p>Effect of initial pH on MO removal yield (<span class="html-italic">C</span><sub>0</sub> = 10 ppm, m = 0.25 g/L, T = 24 °C).</p>
Full article ">Figure 6
<p>Effect of the adsorbent mass of the PAC on the MO elimination rate (<span class="html-italic">C</span><sub>0</sub> = 10 ppm, T = 24 °C).</p>
Full article ">Figure 7
<p>Effect of initial MO dye concentration on efficiency E% of MO removal (m<sub>PAC</sub> = 0.25 g, t = 5–120 min, T = 24 °C).</p>
Full article ">Figure 8
<p>Evolution of the adsorption efficiency of the MO as a function of time for the different adsorbents: SCGs, CAC and PAC (<span class="html-italic">C</span><sub>0</sub> = 10 ppm, m = 0.25 g, T = 19 °C).</p>
Full article ">Figure 9
<p>Adsorption isotherm curve for MO on prepared activated carbon.</p>
Full article ">Figure 10
<p>Effect of CAP carbon mass (<b>a</b>) and (<b>b</b>) solution pH on HA acid removal rate (<span class="html-italic">C</span><sub>0</sub> = 10 ppm, <span class="html-italic">t</span> = 90 min).</p>
Full article ">Figure 11
<p>Evolution of humic acid adsorption yield versus time for different adsorbents: SCGs, CAC and PAC (<span class="html-italic">C</span><sub>0</sub> = 10 ppm, m = 0.25 g, T = 20 °C).</p>
Full article ">Figure 12
<p>Pseudo-first-order kinetic variation for the HA for the adsorption by the PAC.</p>
Full article ">Figure 13
<p>Pseudo-second-order kinetic variation for the adsorption of the HA by the PAC.</p>
Full article ">
Back to TopTop