[go: up one dir, main page]

Next Issue
Volume 14, September
Previous Issue
Volume 14, July
 
 

Catalysts, Volume 14, Issue 8 (August 2024) – 85 articles

Cover Story (view full-size image): Aging is one of the key steps in the preparation of highly active Cu/ZnO-based catalysts for use in the production of methanol. An initially amorphous precipitate transforms into the crystalline precursor phase of zincian malachite, which is characterized by the periodic arrangement of Cu and Zn atoms and has proven advantageous for the quality of the final catalyst. However, aging generally takes between 30 minutes and multiple hours. In this study, we show that aging can be accelerated by more than 90 % when the freshly precipitated suspension is seeded with already aged zincian malachite crystals. Seeding did not alter the physicochemical properties of the aged precursor. Consequently, the catalyst performance in the synthesis of methanol from CO2, as well as from a CO/CO2 mixture, was identical to a catalyst from an unseeded preparation. View this paper
  • Issues are regarded as officially published after their release is announced to the table of contents alert mailing list.
  • You may sign up for e-mail alerts to receive table of contents of newly released issues.
  • PDF is the official format for papers published in both, html and pdf forms. To view the papers in pdf format, click on the "PDF Full-text" link, and use the free Adobe Reader to open them.
Order results
Result details
Section
Select all
Export citation of selected articles as:
15 pages, 1246 KiB  
Article
Biodiesel Production from Waste Frying Oil (WFO) Using a Biomass Ash-Based Catalyst
by Benjamín Nahuelcura, María Eugenia González, Nicolas Gutierrez, Jaime Ñanculeo and Juan Miguel Romero-García
Catalysts 2024, 14(8), 553; https://doi.org/10.3390/catal14080553 - 22 Aug 2024
Viewed by 636
Abstract
Biodiesel, an eco-friendly alternative to conventional fossil fuels, offers reduced emissions like carbon dioxide, sulfur oxides, and soot. This study explores biodiesel production from a blend of waste oils using a novel biomass-based catalyst derived from the bottom ash of a biomass boiler. [...] Read more.
Biodiesel, an eco-friendly alternative to conventional fossil fuels, offers reduced emissions like carbon dioxide, sulfur oxides, and soot. This study explores biodiesel production from a blend of waste oils using a novel biomass-based catalyst derived from the bottom ash of a biomass boiler. Catalyst synthesis involved wet impregnation, a unique approach using previously unreported bottom ash. Characterization via SEM-EDS, BET, FTIR, and XRD revealed its composition and structure. Optimization of biodiesel production involved assessing alcohol molar ratio, catalyst concentration, and reaction time, achieving a maximum FAME concentration of 95% under specific conditions. Blending residual palm oil with waste frying oil enhanced biodiesel properties, demonstrating a maximum FAME concentration at specific catalyst concentration (8%), molar ratio (1:10), and reaction time (2 h). Catalyst reusability, up to three cycles without significant yield variation, showcased its sustainability. The catalyst, primarily composed of calcium, a characteristic biomass bottom ash component, exhibited mesoporous features. Impregnation with eggshells not only altered composition but also ensured a uniform particle size distribution. FTIR and XRD analyses indicated calcium in hydroxide and crystallized forms. Effective catalyst separation methods included decanting or water washing, with optimal biodiesel purity achieved through 3% phosphoric acid washing at 60 °C. Various recovery methods were assessed, highlighting hexane washing as the most efficient, enabling up to three catalyst reuse cycles without substantial efficiency loss. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Particle size distribution fit for RBA and BAC samples. A noticeable difference in variance is observed, which indicates a more homogeneous particle size in the prepared catalyst.</p>
Full article ">Figure 2
<p>FTIR spectra for raw bottom ash RBA, calcined bottom ash CBA, and catalyst (ash/CaO BAC) samples.</p>
Full article ">Figure 3
<p>XRD spectra for raw bottom ash RBA, calcined bottom ash CBA, and catalyst (ash/CaO BAC) samples.</p>
Full article ">Figure 4
<p>Contourplots for the factors molar ratio (MR), catalyst concentration (CW), and reaction time (RT). The response is optimal for low reaction times and for medium values for the others.</p>
Full article ">Figure 5
<p>Yield of FAME according to number of cycles of use for different catalyst recovery methods.</p>
Full article ">Figure 6
<p>FTIR spectrum for catalyst samples after 3 cycles of use. Signal peaks associated with CH<sub>2</sub> groups derived from oil residues are observed.</p>
Full article ">
40 pages, 4101 KiB  
Review
Properties, Industrial Applications and Future Perspectives of Catalytic Materials Based on Nickel and Alumina: A Critical Review
by Guido Busca, Elena Spennati, Paola Riani and Gabriella Garbarino
Catalysts 2024, 14(8), 552; https://doi.org/10.3390/catal14080552 - 22 Aug 2024
Viewed by 633
Abstract
The bulk and surface properties of materials based on nickel and aluminum oxides and hydroxides, as such or after reduction processes, are reviewed and discussed critically. The actual and potential industrial applications of these materials, both in reducing conditions and in oxidizing conditions, [...] Read more.
The bulk and surface properties of materials based on nickel and aluminum oxides and hydroxides, as such or after reduction processes, are reviewed and discussed critically. The actual and potential industrial applications of these materials, both in reducing conditions and in oxidizing conditions, are summarized. Mechanisms for reactant molecule activation are also discussed. Full article
(This article belongs to the Special Issue Feature Papers in "Industrial Catalysis" Section)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of γ-Al<sub>2</sub>O<sub>3</sub> and 20 wt% and 50 wt% loaded NiO/Al<sub>2</sub>O<sub>3</sub>, prepared by impregnation. Arrows are the positions of the peaks of NiO (bunsenite). CoKα radiation.</p>
Full article ">Figure 2
<p>FTIR skeletal spectra (KBr pressed disks) of γ-Al<sub>2</sub>O<sub>3</sub> and 5 wt%, 20 wt% and 50 wt% loaded NiO/Al<sub>2</sub>O<sub>3</sub>, prepared by impregnation, as compared with those of NiAl<sub>2</sub>O<sub>4</sub> and NiO (bunsenite).</p>
Full article ">Figure 3
<p>Visible and near infrared spectra of NiO, NiAl<sub>2</sub>O<sub>4</sub> and 20 wt% NiO/Al<sub>2</sub>O<sub>3</sub> “monolayer” catalyst.</p>
Full article ">Figure 4
<p>Field Emission Scanning Electron Microscopy image of bulk nickel (ex-NiO) after CO<sub>2</sub> hydrogenation experiment. Carbonaceous matter is evident, mainly encapsulating the brighter Ni metal particles.</p>
Full article ">Figure 5
<p>Field Emission Scanning Electron Microscopy image of a Ni/Al<sub>2</sub>O<sub>3</sub>-based catalyst (16% Ni) after ethanol steam reforming experiment (right). Carbonaceous matter is evident, mainly carbon nanotubes that contain some bright nickel particles in the interior.</p>
Full article ">Figure 6
<p>Typical H<sub>2</sub>-TPR curves for samples in the NiO-Al<sub>2</sub>O<sub>3</sub> system.</p>
Full article ">Figure 7
<p>IR spectra of low-temperature adsorption of CO on patterns of γ-Al<sub>2</sub>O<sub>3</sub> (top) and 20 wt% NiO/Al<sub>2</sub>O<sub>3</sub>, prepared by impregnation, oxidized (middle) and reduced (below). Bands at 2211, 2202 and 2188–2195 cm<sup>−1</sup> are due to CO interacting with Al<sup>3+</sup> ions. The band at 2183–2195 cm<sup>−1</sup> is due to CO interacting with Ni<sup>2+</sup>. The bands at 2129, 2059, 2044 and 2020 cm<sup>−1</sup> are due to polycarbonyls of dispersed Ni<sup>0</sup>. The band shifting from 2098 cm<sup>−1</sup> to lower frequencies by decreasing coverage is due to terminal monocarbonyls over Ni metal particle surfaces.</p>
Full article ">Figure 8
<p>FE-SEM micrographs of spent Ni/Al<sub>2</sub>O<sub>3</sub> catalysts after prereduction and CO<sub>2</sub> methanation experiments, recorded using backscattered electrons. Left: 16% Ni (20 wt% NiO) loading; right: 39% (50 wt% NiO) Ni loading. Bright particles are Ni metal particles.</p>
Full article ">
23 pages, 5014 KiB  
Article
Design and Performance of CuNi-rGO and Ag-CuNi-rGO Composite Electrodes for Use in Fuel Cells
by Mohamed Shaban, Aya Mohamed, Mohamed G. M. Kordy, Hamad AlMohamadi, M. F. Eissa and Hany Hamdy
Catalysts 2024, 14(8), 551; https://doi.org/10.3390/catal14080551 - 22 Aug 2024
Viewed by 474
Abstract
This work developed new electrocatalysts for direct alcohol oxidation fuel cells (DAFCs) by using graphene and reduced graphene oxides (GO and rGO) as supporting nanomaterials for copper–nickel (CuNi) nanocomposites. The manufacture of CuNi, CuNi-GO, and CuNi-rGO nanocomposites was realized through the adaptation of [...] Read more.
This work developed new electrocatalysts for direct alcohol oxidation fuel cells (DAFCs) by using graphene and reduced graphene oxides (GO and rGO) as supporting nanomaterials for copper–nickel (CuNi) nanocomposites. The manufacture of CuNi, CuNi-GO, and CuNi-rGO nanocomposites was realized through the adaptation of Hummer’s method and hydrothermal techniques, with subsequent analysis using a range of analytical tools. The electrocatalytic behavior of these materials in DAFCs, with methanol and ethanol as the fuels, was scrutinized through various methods, including cyclic voltammetry, linear sweep, chronoamperometry, and electrochemical impedance spectroscopy. This investigation also assessed the stability and charge transfer dynamics. The rGO-based CuNi nanocomposite demonstrated a remarkable performance boost, showing increases of approximately 319.6% for methanol and 252.6% for ethanol oxidation compared to bare CuNi. The integration of silver nanoparticles into the Ag-CuNi-rGO electrode led to a current density surge to 679.3 mA/g, which signifies enhancements of 254.2% and 812.6% relative to the CuNi-rGO and CuNi electrodes, respectively. These enhancements are ascribed to the augmented densities of hot sites and the synergistic interactions within the nanocatalysts. The findings underscore the potential of Ag and rGO as effective supports for CuNi nanocomposites, amplifying their catalytic efficiency in DAFC applications. Full article
(This article belongs to the Section Nanostructured Catalysts)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) SEM depictions of CuNi NPs and (<b>b</b>) their corresponding nanoparticle size distributions, along with SEM depictions of (<b>c</b>) GO, (<b>d</b>) CuNi-GO, (<b>e</b>) rGO, and (<b>f</b>) CuNi-rGO.</p>
Full article ">Figure 2
<p>FT-IR charts of (<b>a</b>) GO, reduced GO, and CuNi NPs and (<b>b</b>) CuNi-GO and CuNi-rGO composites; (<b>c</b>) Raman spectra of CuNi-GO and CuNi-rGO nanocomposites; and (<b>d</b>) XRD patterns of CuNi, CuNi-GO, and CuNi-rGO nanocomposites.</p>
Full article ">Figure 3
<p>(<b>a</b>) Absorbance behaviors and (<b>b</b>) bandgap estimation for rGO, CuNi, CuNi-GO, and CuNi-rGO.</p>
Full article ">Figure 4
<p>The influence of sample compositions on the electrocatalytic efficacy in 1 M KOH solutions at 20 °C, targeting the oxidation of (<b>a</b>–<b>c</b>) 0.5 M ethanol and (<b>d</b>–<b>f</b>) 2 M methanol using CuNi, CuNi-GO, CuNi-rGO, and Ag-CuNi-rGO.</p>
Full article ">Figure 5
<p>The influence of (<b>a</b>) ethanol and (<b>b</b>) methanol concentrations on the electrocatalytic CuNi-rGO performance at 20 °C, 100 mV/s.</p>
Full article ">Figure 6
<p>Scanning rate effects on the electrocatalytic efficacy of (<b>a</b>) CuNi-rGO in 0.5 M ethanol and (<b>b</b>) CuNi-rGO and (<b>c</b>) Ag-CuNi-rGO in 2 M methanol at 20 °C.</p>
Full article ">Figure 7
<p>LSV responses at 100 mV.s<sup>−1</sup> for CuNi, CuNi-GO, CuNi-rGO, and Ag-CuNi-rGO in (<b>a</b>) ethanol and (<b>b</b>) methanol, (<b>c</b>) the alterations in current density over time, as determined by CAM analysis, in ethanol and methanol.</p>
Full article ">Figure 8
<p>(<b>a</b>) Nyquist and (<b>b</b>) Bode diagrams for the electrodes Ag-CuNi-rGO, CuNi-rGO, CuNi-GO, and CuNi when tested in a 2 M methanol solution.</p>
Full article ">
12 pages, 7415 KiB  
Article
Photoinduced Mechanisms of C–S Borylation of Methyl(p-tolyl)Sulfane with Bis(Pinacolato)diboron: A Density Functional Theory Investigation
by Yuxiao Ming, Tiantian Feng, Bin Chen and Dagang Zhou
Catalysts 2024, 14(8), 550; https://doi.org/10.3390/catal14080550 - 22 Aug 2024
Viewed by 331
Abstract
The reaction mechanisms of C–S borylation of aryl sulfides catalyzed with 1,4-benzoquinone (BQ) were investigated by employing the M06-2X-D3/ma-def2-SVP method and basis set. In this study, the SMD model was taken to simulate the solvent effect of 1,4-dioxane. Also, TD-DFT calculations of BQ [...] Read more.
The reaction mechanisms of C–S borylation of aryl sulfides catalyzed with 1,4-benzoquinone (BQ) were investigated by employing the M06-2X-D3/ma-def2-SVP method and basis set. In this study, the SMD model was taken to simulate the solvent effect of 1,4-dioxane. Also, TD-DFT calculations of BQ and methyl(p-tolyl)sulfane were performed in an SMD solvent model. The computational results indicated that BQ and methyl(p-tolyl)sulfane, serving as a photo-catalyst, would be excited under a blue LED of 450 nm, aligning well with experimental observations. Additionally, the role of 3O2 was investigated, revealing that it could be activated into 1O2 from the released energy of 1[BQ + methyl(p-tolyl)sulfane]* or 3[BQ + methyl(p-tolyl)sulfane]*→BQ + methyl(p-tolyl)sulfane process. Then, 1O2, bis(pinacolato)diboron, and methyl(p-tolyl)sulfane would, through a series of reactions, yield the final product, P. The Gibbs free energy surface shows that path a2-2 is optimal, and this path has fewer steps and a lower energy barrier. Electron spin density isosurface graphs were employed to analyze the structures and elucidate the single electron distribution. These computational results offer valuable insights into the studied interactions and related processes and shed light on the mechanisms governing C–S borylation from aryl sulfides and b2pin2 catalyzed with BQ and methyl(p-tolyl)sulfane. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Frontier molecular orbital of BQ and R1 + BQ models.</p>
Full article ">Figure 2
<p>The ρ<sub>hole</sub> (blue) and ρ<sub>ele</sub> (green) of S<sub>0</sub>→S<sub>1</sub>, S<sub>0</sub>→S<sub>2</sub>, and S<sub>0</sub>→S<sub>3</sub> of R1 + BQ.</p>
Full article ">Figure 3
<p>The Gibbs free energy surfaces in photocatalysis process.</p>
Full article ">Figure 4
<p>The Gibbs free energy surfaces of Paths a1-1, a1-2, and a1-3.</p>
Full article ">Figure 5
<p>The electron spin density isosurface graphs of some structures in Path a1 and a2.</p>
Full article ">Figure 6
<p>The Gibbs free energy surfaces from IM1 to P via probable Path a2-1.</p>
Full article ">Figure 7
<p>The Gibbs free energy surfaces from IM1 to P via probable Path a2-2.</p>
Full article ">Scheme 1
<p>C–S borylation of aryl sulfides.</p>
Full article ">Scheme 2
<p>The total reaction from R1 + R2→P.</p>
Full article ">Scheme 3
<p>The detailed reaction photocatalysis process.</p>
Full article ">Scheme 4
<p>The detailed reaction process from IM1 to P via two probable Paths, a2-1 and a2-2.</p>
Full article ">
14 pages, 6272 KiB  
Article
Biodiesel Synthesis from Coconut Oil Using the Ash of Citrus limetta Peels as a Renewable Heterogeneous Catalyst
by Priyal Kaushik, Gurmeet Kaur and Imran Hasan
Catalysts 2024, 14(8), 549; https://doi.org/10.3390/catal14080549 - 22 Aug 2024
Viewed by 567
Abstract
The synthesis of biodiesel can be achieved using either homogeneous or heterogeneous catalysts. Given the non-renewable nature of homogeneous catalysts, heterogeneous catalysts are increasingly preferred for biodiesel production. Agricultural wastes serve as a viable source for these heterogeneous catalysts, contributing to environmental sustainability. [...] Read more.
The synthesis of biodiesel can be achieved using either homogeneous or heterogeneous catalysts. Given the non-renewable nature of homogeneous catalysts, heterogeneous catalysts are increasingly preferred for biodiesel production. Agricultural wastes serve as a viable source for these heterogeneous catalysts, contributing to environmental sustainability. This study introduces a novel, eco-friendly, cost-effective, and efficient heterogeneous catalyst that was developed and derived from Citrus limetta peels for biodiesel production. The catalyst was thoroughly characterized using Fourier-transform infrared spectroscopy (FTIR), X-ray diffractograms (XRD), Field Emission Scanning electron microscopy (FESEM), and energy-dispersive X-ray (EDX). Coconut oil, a rich and renewable resource, was used as the feedstock for the biodiesel synthesis. The conversion process was confirmed by 1H NMR, IR spectra, mass spectra, and 13C NMR of the biodiesel. The developed method using the Citrus limetta peel-derived catalyst demonstrated a 100% yield. The results show the optimum conditions for biodiesel synthesis are 1 w/v (for the catalytical dose), with a 6:1 methanol/oil ratio at 60 °C for 3 h. The synthesized biodiesel exhibited a high cetane value of 54, contributing to improved ignition and reduced engine noise. Its sulfur-free composition, boiling point of 294 °C, high viscosity of 2.5 mm2/s, acid value of 0.09 mgKOH/g, and flash point of 142 °C enhance its environmental profile. The catalyst can be used for up to five cycles, underscoring its potential as a cost-effective and sustainable approach for biodiesel production. Full article
(This article belongs to the Special Issue Novel Materials for Heterogeneous Catalysis and Energy Conversion)
Show Figures

Figure 1

Figure 1
<p>Biodiesel synthesis using different waste biomass in literature.</p>
Full article ">Figure 2
<p>Transesterification reaction for biodiesel synthesis.</p>
Full article ">Figure 3
<p>Catalyst dose effect on FAME conversion.</p>
Full article ">Figure 4
<p>Methanol/oil effect on FAME conversion.</p>
Full article ">Figure 5
<p>Time effect on FAME conversion.</p>
Full article ">Figure 6
<p>Several cycles affect FAME conversion.</p>
Full article ">Figure 7
<p>XRD data of (<b>a</b>) fresh catalyst and (<b>b</b>) catalyst after multiple cycles.</p>
Full article ">Figure 8
<p>FESEM data of CCLP.</p>
Full article ">Figure 8 Cont.
<p>FESEM data of CCLP.</p>
Full article ">Figure 9
<p>EDS of the catalyst.</p>
Full article ">Figure 10
<p>IR data of (<b>a</b>) fresh catalyst, (<b>b</b>) catalyst after leaching, and (<b>c</b>) synthesized biodiesel.</p>
Full article ">Figure 11
<p>Lauric acid is present in coconut oil.</p>
Full article ">Figure 12
<p>Mass spectra of synthesized biodiesel.</p>
Full article ">Figure 13
<p><sup>1</sup>H NMR data of biodiesel.</p>
Full article ">Figure 14
<p><sup>13</sup>C NMR data of synthesized biodiesel.</p>
Full article ">Figure 15
<p>Metals present in CCLP.</p>
Full article ">Figure 16
<p>Metal interaction with triglyceride.</p>
Full article ">Figure 17
<p>The enzymatic action of the catalyst.</p>
Full article ">
14 pages, 4486 KiB  
Article
Efficient and Robust Photodegradation of Dichlorvos Pesticide by BiOBr/WO2.72 Nanocomposites with Type-I Heterojunction under Visible Light Irradiation
by Aoyun Meng, Wen Li, Zhen Li and Jinfeng Zhang
Catalysts 2024, 14(8), 548; https://doi.org/10.3390/catal14080548 - 21 Aug 2024
Viewed by 463
Abstract
In this study, we developed novel BiOBr/WO2.72 nanocomposites (abbreviated as BO/WO) and systematically investigated their photocatalytic degradation performance against the pesticide dichlorvos under visible light irradiation. The experimental results demonstrated that the BO/WO nanocomposites achieved an 85.4% degradation of dichlorvos within 80 [...] Read more.
In this study, we developed novel BiOBr/WO2.72 nanocomposites (abbreviated as BO/WO) and systematically investigated their photocatalytic degradation performance against the pesticide dichlorvos under visible light irradiation. The experimental results demonstrated that the BO/WO nanocomposites achieved an 85.4% degradation of dichlorvos within 80 min. In comparison, the BO alone achieved a degradation degree of 66.8%, and the WO achieved a degradation degree of 64.7%. Furthermore, the BO/WO nanocomposites retained 96% of their initial activity over five consecutive cycles, demonstrating exceptional stability. Advanced characterization techniques, such as high-resolution transmission electron microscopy (HRTEM), X-ray photoelectron spectroscopy (XPS), and electron paramagnetic resonance (EPR) confirmed the composition and catalytic mechanism of the composite material. The findings indicated that the BO/WO nanocomposites, through their optimized Type-I heterojunction structure, achieved efficient separation and transport of photogenerated electron–hole pairs, significantly enhancing the degree of degradation of organophosphate pesticides. This research not only propels the development of high-performance photocatalytic materials, but also provides innovative strategies and a robust scientific foundation for mitigating global organophosphate pesticide pollution, underscoring its substantial potential for environmental remediation. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>a</b>) BO and, BO/WO nanocomposites, and WO; (<b>b</b>) FT-IR spectra of BO, BO/WO nanocomposites, and WO.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) BO, (<b>b</b>) WO, and (<b>c</b>) BO/WO; EDS spectra of (<b>d</b>) BO, (<b>e</b>) WO, and (<b>f</b>) BO/WO. TEM images of (<b>g</b>) BO, (<b>h</b>) WO, and (<b>i</b>) BO/WO (inset shows the HRTEM image of the BO/WO nanocomposites).</p>
Full article ">Figure 3
<p>XPS spectra: (<b>a</b>) Survey spectra; (<b>b</b>) Br 3d; (<b>c</b>) O 1s; (<b>d</b>) Bi 4f; (<b>e</b>) W 4f; (<b>f</b>) BET surface area of catalysts.</p>
Full article ">Figure 4
<p>(<b>a</b>) Photocatalytic dichlorvos degradation profiles of BO, (<b>b</b>) BO/WO nanocomposites, and (<b>c</b>) WO under visible light irradiation. Bottom panels show the corresponding degradation extents over time.</p>
Full article ">Figure 5
<p>Photocatalytic stability of (<b>a</b>) BO and, (<b>b</b>) BO/WO nanocomposites, and of (<b>c</b>) WO for the degradation of dichlorvos under visible light irradiation across multiple cycles.</p>
Full article ">Figure 6
<p>Degradation degree of BO/WO with (<b>a</b>) IPA, (<b>b</b>) ascorbic acid, (<b>c</b>) TEOA, and (<b>d</b>) KBrO<sub>3</sub>.</p>
Full article ">Figure 7
<p>(<b>a</b>) UV-vis absorption spectra; (<b>b</b>) Tauc plot for WO; (<b>c</b>) Tauc plot for BO; (<b>d</b>) EPR spectra of BO/WO; (<b>e</b>) electrostatic potential of WO; (<b>f</b>) electrostatic potential of BO.</p>
Full article ">Figure 8
<p>(<b>a</b>) Photocurrent responses of the BO, WO, and BO/WO nanocomposites; (<b>b</b>) EIS Nyquist plots of the BO, WO, and BO/WO nanocomposites.</p>
Full article ">Figure 9
<p>The photocatalysis mechanism of the BO/WO nanocomposites under visible light.</p>
Full article ">Scheme 1
<p>Schematic illustration of the synthesis of BO/WO nanocomposites.</p>
Full article ">
6 pages, 1865 KiB  
Editorial
Fifty Years of Research in Environmental Photocatalysis: Scientific Advances, Discoveries, and New Perspectives
by Chantal Guillard and Didier Robert
Catalysts 2024, 14(8), 547; https://doi.org/10.3390/catal14080547 - 21 Aug 2024
Viewed by 630
Abstract
The concept of “photocatalysis” was discovered at the beginning of the twentieth century, and research in this area has especially intensified over the past 50 years [...] Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the principles of photocatalysis.</p>
Full article ">Figure 2
<p>Photocatalysis applications (adapted from [<a href="#B11-catalysts-14-00547" class="html-bibr">11</a>]).</p>
Full article ">Figure 3
<p>Number of publications published in 2023 in the field of photocatalysis (Scopus source).</p>
Full article ">Figure 4
<p>Number of publications published in 2023 (<b>a</b>) on environmental photocatalysis and (<b>b</b>) on the use of photocatalysis for energy (Scopus source).</p>
Full article ">Figure 4 Cont.
<p>Number of publications published in 2023 (<b>a</b>) on environmental photocatalysis and (<b>b</b>) on the use of photocatalysis for energy (Scopus source).</p>
Full article ">
12 pages, 2695 KiB  
Article
Impact of Oxygen-Containing Groups on Pd/C in the Catalytic Hydrogenation of Acetophenone and Phenylacetylene
by Pengyao You, Liming Wu, Lu Zhou, Yong Xu and Ruixuan Qin
Catalysts 2024, 14(8), 545; https://doi.org/10.3390/catal14080545 - 21 Aug 2024
Viewed by 552
Abstract
Pd/C catalysts play a pivotal role in contemporary chemical industries due to their exceptional performance in diverse hydrogenation processes and organic reactions. Over the past century, researchers have extensively explored the factors influencing Pd/C catalyst performance, particularly emphasizing the impact of oxygen-containing groups [...] Read more.
Pd/C catalysts play a pivotal role in contemporary chemical industries due to their exceptional performance in diverse hydrogenation processes and organic reactions. Over the past century, researchers have extensively explored the factors influencing Pd/C catalyst performance, particularly emphasizing the impact of oxygen-containing groups through oxidation or reduction modifications. However, most studies use respective Pd/C catalysts to analyze the catalytic reactions of one or a class of chemical bonds (polar or non-polar). This study investigates alterations in Pd/C catalysts during temperature-programmed reduction (TPR) and evaluates the hydrogenation activity of unsaturated polar bonds (C=O, acetophenone) and non-polar bonds (C≡C, phenylacetylene) in Pd/C catalysts. The experimental results indicate that the reduction of Pd/C decreases the content of oxygen-containing groups, reducing hydrogenation activity for acetophenone but increasing it for phenylacetylene. This research highlights the preference of regular Pd surfaces for non-polar bond reactions and the role of Pd/oxide sites in facilitating polar bond hydrogenation. These discoveries offer essential insights into how oxygen-containing groups influence catalytic performance and allow us to propose potential avenues for enhancing the design and production of Pd/C catalysis. Full article
(This article belongs to the Special Issue Heterogeneous Catalysis for Selective Hydrogenation)
Show Figures

Figure 1

Figure 1
<p>The TPR-MS signals of the Pd/C catalyst recorded in 5% H<sub>2</sub>/Ar.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>c</b>) TEM images and (<b>d</b>–<b>f</b>) particle size distribution statistics for 60 °C-H<sub>2</sub>-Pd/C, 500 °C-TPR-Pd/C, and 800 °C-TPR-Pd/C, respectively.</p>
Full article ">Figure 3
<p>(<b>a</b>) XRD of initial Pd/C, 60 °C-H<sub>2</sub>-Pd/C, 500 °C-TPR-Pd/C, and 800 °C-TPR-Pd/C. (<b>b</b>) CO-DRIFTS of 60 °C-H<sub>2</sub>-Pd/C and 500 °C-TPR-Pd/C. Prior to CO-DRIFTS collection, the sample was purged successively with Ar, 5% CO/Ar, and Ar for 10 min each.</p>
Full article ">Figure 4
<p>Mathematical relationship between particle diameter and activity of a regular polyhedron. (<b>a</b>) FCC, regular tetrahedron; (<b>b</b>) FCC, regular octahedron; (<b>c</b>) FCC, cube; (<b>d</b>) BCC, cube. FCC: Face Center Cubic; BCC: Body Center Cubic.</p>
Full article ">Figure 5
<p>Catalytic performance of 60 °C-H<sub>2</sub>-Pd/C, 500 °C-TPR-Pd/C, and 800 °C-TPR-Pd/C. Hydrogenation of acetophenone: (<b>a</b>) catalytic activity, (<b>b</b>) turnover frequency (TOF), and (<b>c</b>) analysis of activity related to particle size. Hydrogenation of phenylacetylene: (<b>d</b>) catalytic activity, (<b>e</b>) TOF, and (<b>f</b>) analysis of activity related to particle size. Reaction conditions: 60 °C, 1 bar H<sub>2</sub>, 5 mL ethanol, n(acetophenone):n(Pd) = 500:1; n(phenylacetylene):n(Pd) = 2000:1. k = TOF = (the molar of substrates conversion)/(the molar of Pd on the surface)/(conversion time).</p>
Full article ">
14 pages, 4821 KiB  
Article
Research on Cu-Site Modification of g-C3N4/CeO2-like Z-Scheme Heterojunction for Enhancing CO2 Reduction and Mechanism Insight
by Yiying Zhou, Junxi Cai, Yuming Sun, Shuhan Jia, Zhonghuan Liu, Xu Tang, Bo Hu, Yue Zhang, Yan Yan and Zhi Zhu
Catalysts 2024, 14(8), 546; https://doi.org/10.3390/catal14080546 - 20 Aug 2024
Viewed by 373
Abstract
In this work, the successful synthesis of a Cu@g-C3N4/CeO2-like Z-scheme heterojunction through hydrothermal and photo-deposition methods represents high CO2 reduction activity with remarkable CO selectivity, as evidenced by the impressive CO yield of 33.8 [...] Read more.
In this work, the successful synthesis of a Cu@g-C3N4/CeO2-like Z-scheme heterojunction through hydrothermal and photo-deposition methods represents high CO2 reduction activity with remarkable CO selectivity, as evidenced by the impressive CO yield of 33.8 μmol/g for Cu@g-C3N4/CeO2, which is over 10 times higher than that of g-C3N4 and CeO2 individually. The characterization and control experimental results indicate that the formation of heterojunctions and the introduction of Cu sites promote charge separation and the transfer of hot electrons, as well as the photothermal effect, which are the essential reasons for the improved CO2 reduction activity. Remarkably, Cu@g-C3N4/CeO2 still exhibits about 92% performance even after multiple cycles. In situ FTIR was utilized to confirm the production of COOH* at 1472 cm−1 and to elucidate the mechanism behind the high selectivity for CO production. The study’s investigation into the wide-ranging applicability of the Cu@g-C3N4/CeO2-like Z-scheme heterojunction catalysts is noteworthy, and the exploration of potential reaction mechanisms for CO2 reduction adds valuable insights to the field of catalysis. Full article
(This article belongs to the Special Issue Mineral-Based Composite Catalytic Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structures of monolayer CeO<sub>2</sub> (<b>a</b>), <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> (<b>b</b>), Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>c</b>).</p>
Full article ">Figure 2
<p>TEM images of the <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> (<b>a</b>), CeO<sub>2</sub> (<b>b</b>), Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>c</b>), SEM images of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> (<b>d</b>), CeO<sub>2</sub> (<b>e</b>), Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>f</b>), EDX mapping of Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>g1</b>–<b>g5</b>).</p>
Full article ">Figure 3
<p>XRD patterns (<b>a</b>), XPS survey spectra (<b>b</b>) of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, CeO<sub>2</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub>, high resolution of C1s of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub>(<b>c</b>), O1s XPS spectra of CeO<sub>2</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>d</b>), N1s XPS spectra of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>e</b>), Ce 3d XPS spectra of CeO<sub>2</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> (<b>f</b>).</p>
Full article ">Figure 4
<p>PL spectra (<b>a</b>), FL spectra (<b>b</b>), UV–Vis DRS (<b>c</b>), transient photocurrent response (<b>d</b>) and EIS Nyquist plots (<b>e</b>) of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and CeO<sub>2</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub>, Mott–Schottky (<b>f</b>) of Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub>.</p>
Full article ">Figure 5
<p>CO yields with condensed water ((<b>a</b>), Xenon lamp, 1000 mW/cm<sup>2</sup>, 6 °C); the CO yields over <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and CeO<sub>2</sub> and Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> hybrids without condensed water ((<b>b</b>), Xenon lamp, 1000 mW/cm<sup>2</sup>); the recycling test for CO<sub>2</sub> reduction using Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> without condensed water ((<b>c</b>), Xenon lamp, 1000 mW/cm<sup>2</sup>).</p>
Full article ">Figure 6
<p>(<b>a</b>) XRD comparison of photocatalysis before and after photocatalysis; (<b>b</b>) Evolution of CO after 4 h of reaction under various reaction conditions. In the case of N<sub>2</sub>, no light, normal and without photocatalyst, respectively.</p>
Full article ">Figure 7
<p>In situ FTIR spectra of CO<sub>2</sub> adsorption (<b>a</b>) and reaction (<b>b</b>) of Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> collected at different time intervals.</p>
Full article ">Figure 8
<p>Schematic diagram of the mechanism of CO<sub>2</sub> reduction by Cu@<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/CeO<sub>2</sub> under visible light irradiation.</p>
Full article ">
18 pages, 3457 KiB  
Article
Influence of UV-A Light Modulation on Phenol Mineralization by TiO2 Photocatalytic Process Coadjuvated with H2O2
by Nicola Morante, Luca De Guglielmo, Nunzio Oliva, Katia Monzillo, Nicola Femia, Giulia Di Capua, Vincenzo Vaiano and Diana Sannino
Catalysts 2024, 14(8), 544; https://doi.org/10.3390/catal14080544 - 20 Aug 2024
Viewed by 656
Abstract
This work examined the influence of UV-A light modulation on the photocatalytic process coadjuvated with H2O2 to mineralize phenol in an aqueous solution. A fixed-bed batch photocatalytic reactor with a flat-plate geometry, irradiated by UV-A LEDs, was employed. The successful [...] Read more.
This work examined the influence of UV-A light modulation on the photocatalytic process coadjuvated with H2O2 to mineralize phenol in an aqueous solution. A fixed-bed batch photocatalytic reactor with a flat-plate geometry, irradiated by UV-A LEDs, was employed. The successful deposition of commercial TiO2 PC105 on a steel plate (SP) was achieved, and the structured photocatalyst was characterized using Raman spectroscopy, specific surface area (SSA) measurements, and UV–vis DRS analysis. These analyses confirmed the formation of a titania coating in the anatase phase with a bandgap energy of 3.25 eV. Various LED-dimming techniques, with both fixed and variable duty cycle values, were tested to evaluate the stability of the photocatalyst’s activity and the influence of operating parameters during the mineralization of 450 mL of a phenol solution. The optimal operating parameters were identified as an initial phenol concentration of 10 ppm, a hydrogen peroxide dosage of 0.208 g L−1, and triangular variable duty cycle light modulation. Under these conditions, the highest apparent phenol degradation kinetic constant (0.39 min−1) and the total mineralization were achieved. Finally, the energy consumption for mineralizing 90% phenol in one cubic meter of treated water was determined, showing the greatest energy savings with triangular light modulation. Full article
(This article belongs to the Special Issue Commemorative Special Issue for Prof. Dr. Dion Dionysiou)
Show Figures

Figure 1

Figure 1
<p>Raman spectra of PC105/SP, PC105, and SP samples.</p>
Full article ">Figure 2
<p>UV–vis diffuse reflectance spectra of PC105/SP and SP samples.</p>
Full article ">Figure 3
<p>Tauc plot for the energy bandgap estimation of PC105/SP sample.</p>
Full article ">Figure 4
<p>(<b>a</b>) Phenol degradation under UV-A light registered with FD light modulation for five reuse cycles; (<b>b</b>) mineralization efficiency after 180 min of irradiation and apparent kinetic degradation constant values obtained with FD light modulation for five reuse cycles.</p>
Full article ">Figure 5
<p>(<b>a</b>) Phenol degradation under UV-A light registered for different initial phenol concentrations; (<b>b</b>) mineralization efficiency after 180 min of irradiation and apparent kinetic degradation constant values obtained for different initial phenol concentrations.</p>
Full article ">Figure 6
<p>(<b>a</b>) Phenol degradation under UV-A light registered for different reaction conditions; (<b>b</b>) mineralization efficiency after 30 min of irradiation and apparent kinetic degradation constant values obtained for different reaction conditions.</p>
Full article ">Figure 6 Cont.
<p>(<b>a</b>) Phenol degradation under UV-A light registered for different reaction conditions; (<b>b</b>) mineralization efficiency after 30 min of irradiation and apparent kinetic degradation constant values obtained for different reaction conditions.</p>
Full article ">Figure 7
<p>(<b>a</b>) Phenol degradation under UV-A light irradiation registered for different initial H<sub>2</sub>O<sub>2</sub> dosages; (<b>b</b>) mineralization efficiency after 30 min of irradiation and apparent kinetic degradation constant values obtained for different initial H<sub>2</sub>O<sub>2</sub> dosages.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) Phenol degradation under UV-A light irradiation registered for different initial H<sub>2</sub>O<sub>2</sub> dosages; (<b>b</b>) mineralization efficiency after 30 min of irradiation and apparent kinetic degradation constant values obtained for different initial H<sub>2</sub>O<sub>2</sub> dosages.</p>
Full article ">Figure 8
<p>H<sub>2</sub>O<sub>2</sub> consumption after 180 min under UV-A irradiation on PC105/SP by varying initial H<sub>2</sub>O<sub>2</sub> dosage.</p>
Full article ">Figure 9
<p>Mineralization efficiency after 10 min of irradiation and apparent kinetic degradation constant values obtained using different light modulation.</p>
Full article ">Figure 10
<p>(<b>a</b>) Picture of the steel plate before deposition of the photocatalyst; (<b>b</b>) picture of the PC105/SP-structured photocatalyst after deposition of the photocatalyst.</p>
Full article ">Figure 11
<p>Experimental setup scheme.</p>
Full article ">Figure 12
<p>Values of the incident light intensity inside the reactor by varying the feed current to UV-A LEDs.</p>
Full article ">
12 pages, 3035 KiB  
Article
Dual-Enzyme-Cascade Catalysis for PET Biodegradation Based on a Variable-Temperature Program
by Dong Lu, Jinglong Wu, Shuming Jin, Qiuyang Wu, Fang Wang, Li Deng and Kaili Nie
Catalysts 2024, 14(8), 543; https://doi.org/10.3390/catal14080543 - 20 Aug 2024
Viewed by 967
Abstract
As an environmentally friendly technology, enzymatic degradation of waste polyethylene terephthalate (PET) has great application potential. Mono (hydroxyethyl) terephthalate (MHET), an intermediate product of PET degradation, accumulates during the degradation process. MHET reduces the activity of PETase and influences further enzymatic degradation. The [...] Read more.
As an environmentally friendly technology, enzymatic degradation of waste polyethylene terephthalate (PET) has great application potential. Mono (hydroxyethyl) terephthalate (MHET), an intermediate product of PET degradation, accumulates during the degradation process. MHET reduces the activity of PETase and influences further enzymatic degradation. The combined catalysis of MHETase and PETase is an effective strategy to solve this problem. However, the difference in thermostability between MHETase and PETase limits their combination. In our previous study, a PETase of muEst1 exhibited acceptable PET-degradation ability, but the abundant MHET accumulation in its degradation products limited its further application. In this study, MHETases with good thermostability were screened for combination with muEst1 for the cascade reaction of PET degradation, and a two-stage variable-temperature program was developed. The results of this investigation show that this approach results in a PET-degradation rate of 92.71% with a terephthalic acid content above 85.9%. This investigation provides an alternative method for scaled-up enzymatic PET degradation. Full article
(This article belongs to the Section Biocatalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Biodegradation of particles of a post-consumer commercial PET bottle. Product concentrations and degradation rates at 3‰ (<b>A</b>), 3% (<b>B</b>), and 5% (<b>C</b>) muEst1 dosages.</p>
Full article ">Figure 2
<p>SDS-PAGE of different hydrolases: (<b>A</b>) supernatant of the expression cell culture. (<b>B</b>) pellet of the expression cell culture. Lane 1: muEst1, lane 2: <span class="html-italic">KL</span>MHETase, lane 3: <span class="html-italic">Bs</span>MHETase, lane 4: <span class="html-italic">Tt</span>MHETase, lane 5: <span class="html-italic">Is</span>MHETase, lane 6: <span class="html-italic">KL</span>_muEst1, lane 7: <span class="html-italic">Bs</span>_muEst1, lane 8: <span class="html-italic">Tt</span>_muEst1, lane 9: <span class="html-italic">Is</span>-muEst1. The individual and fusion protein sizes were as follows: muEst1 29.4 kDa, <span class="html-italic">KL</span>MHETase 28.5 kDa, <span class="html-italic">Bs</span>MHETase 54.4 kDa, <span class="html-italic">Tt</span>MHETase 26.4 kDa, <span class="html-italic">Is</span>MHETase 63.8 kDa, <span class="html-italic">KL</span>_muEst1 60.1 kDa, <span class="html-italic">Bs</span>_muEst1 86.1 kDa, <span class="html-italic">Tt</span>_muEst1 58.1 kDa, and <span class="html-italic">Is</span>_muEst1 98.5 kDa. <span class="html-italic">Is</span>MHETase--MHETase from <span class="html-italic">Ideonella sakaiensis</span>, <span class="html-italic">KL</span>MHETase--redesigned from thermophilic carboxylesterase Est30 (<span class="html-italic">Geobacillus stearothermophilus</span>), <span class="html-italic">Tt</span>MHETase--MHETase from <span class="html-italic">Thermus thermophilus</span>, and <span class="html-italic">Bs</span>MHETase--MHETase from <span class="html-italic">Bacillus subtilis</span>.</p>
Full article ">Figure 3
<p>The results of PET degradation by combinations of muEst1 and MHETases. (<b>A</b>) PET degradation at 50 °C for 8 h; (<b>B</b>) PET degradation at 65 °C for 8 h; (<b>C</b>) PET degradation at 50 °C for 24 h; (<b>D</b>) PET degradation at 65 °C for 24 h; (<b>E</b>) thermostability of enzymes at 50 °C and 65 °C; (<b>F</b>) comparison of enzymatic activity.</p>
Full article ">Figure 4
<p>PET degradation by fusion proteins.</p>
Full article ">Figure 5
<p>PET degradation by the three-stage process. (<b>A</b>) Scheme of the three-stage process; (<b>B</b>) Comparison of PET degradation by different processes; (<b>C</b>) Analysis of product composition and release.</p>
Full article ">Figure 6
<p>PET degradation in the two-stage process. (<b>A</b>) Scheme of the two-stage process; (<b>B</b>) Product release in the single-enzyme muEst1 samples; (<b>C</b>) Product release in the dual-enzyme samples.</p>
Full article ">Figure 7
<p>PET degradation in a 5 L bioreactor.</p>
Full article ">
14 pages, 3798 KiB  
Article
Oxidation in Flow Using an Ionic Immobilized TEMPO Catalyst on an Ion Exchange Resin
by Johannes Gmeiner and Gerrit A. Luinstra
Catalysts 2024, 14(8), 542; https://doi.org/10.3390/catal14080542 - 19 Aug 2024
Viewed by 452
Abstract
An ionic heterogenized catalyst system for Anelli oxidation has been developed using potassium 4-sulfonato-oxy-2,2,6,6-tetramethylpiperidine-1-yloxyl (TEMPO-4-sulfate) and anion exchange beads as support material. The catalytic oxidation of benzyl alcohol by bis(acetoxy)iodobenzene (BAIB) in acetonitrile with the modified beads gives a 94% yield of benzaldehyde [...] Read more.
An ionic heterogenized catalyst system for Anelli oxidation has been developed using potassium 4-sulfonato-oxy-2,2,6,6-tetramethylpiperidine-1-yloxyl (TEMPO-4-sulfate) and anion exchange beads as support material. The catalytic oxidation of benzyl alcohol by bis(acetoxy)iodobenzene (BAIB) in acetonitrile with the modified beads gives a 94% yield of benzaldehyde within 60 min (batch operation). The beads give about the same conversion of benzyl alcohol in six consecutive cycles when reused after simple washing, albeit with a somewhat longer half-life time. The TEMPO entity could be removed from the beads using a sodium chloride/sodium hydroxy mixture. Reloading the beads with TEMPO-4-sulfate restored about 80% of their initial catalytic action. This exemplifies that the catalytic activity in a fixed bed can be regained without the need for cleaning and repacking the reactor. Preliminary experiments in a fixed bed show that a constant benzyl alcohol conversion of 84% over 10 residence times (as plug flow) can be achieved by the in-flow execution of the oxidation reaction. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Setup for Anelli oxidation in flow: ion exchange resins are loaded with anionic TEMPO-4-sulfate (green), and benzyl alcohol in solution is subsequently oxidized to benzaldehyde (orange); catalyst column is regenerated for TEMPO-reloading with a 10% NaCl/1% NaOH solution (blue).</p>
Full article ">Figure 2
<p>Light microscope images of ROTI<sup>®</sup>Change 1 × 8 20–50 mesh untreated beads with an average diameter of 622 (±99) µm (<b>a</b>) and after loading with TEMPO-4-sulfate with a diameter of 663 (±124) µm (<b>b</b>).</p>
Full article ">Figure 3
<p>Absorption of TEMPO-4 sulfate in UV–Vis spectra on ROTI<sup>®</sup>Change ion exchange resins in batch (black) and flow loading (orange) (c.f. <a href="#app1-catalysts-14-00542" class="html-app">Figure S1</a>).</p>
Full article ">Figure 4
<p>ATR FT-IR spectra (<b>a</b>) and EDX analysis (<b>b</b>) of raw, TEMPO-4-sulfate-loaded, and unloaded ROTI<sup>®</sup>Change ion exchange resins after the sixth oxidation cycle.</p>
Full article ">Figure 5
<p>Oxidation of benzyl alcohol to benzaldehyde in various solvents after 1 h.</p>
Full article ">Figure 6
<p>Oxidation of benzyl alcohol with reuse of the catalyst beads and after unloading and reloading with TEMPO-4-sulfate.</p>
Full article ">Figure 7
<p>SEM images of ROTI<sup>®</sup>Change ion exchange resins in maiden state (raw; <b>left</b> image) and unloaded resin after six oxidation cycles in the batch (<b>right</b>).</p>
Full article ">Figure 8
<p>Oxidation of benzyl alcohol in flow at different flowrates.</p>
Full article ">Scheme 1
<p>TEMPO-4-sulfate exchange process at the alkyl ammonium chloride form of ROTI<sup>®</sup>Change.</p>
Full article ">Scheme 2
<p>Setup for continuous flow oxidation of benzyl alcohol to benzaldehyde using a packed and loaded ion exchange column.</p>
Full article ">
15 pages, 4832 KiB  
Article
Cu/MgO as an Efficient New Catalyst for the Non-Oxidative Dehydrogenation of Ethanol into Acetaldehyde
by Chao Tian, Yinghong Yue, Changxi Miao, Weiming Hua and Zi Gao
Catalysts 2024, 14(8), 541; https://doi.org/10.3390/catal14080541 - 19 Aug 2024
Viewed by 435
Abstract
The non-oxidative dehydrogenation of ethanol into acetaldehyde is one of the efficient solutions for biomass upgrading. In this work, a series of copper catalysts supported on MgO with different Cu loadings ranging from 2.5% to 20% were prepared by an impregnation method. The [...] Read more.
The non-oxidative dehydrogenation of ethanol into acetaldehyde is one of the efficient solutions for biomass upgrading. In this work, a series of copper catalysts supported on MgO with different Cu loadings ranging from 2.5% to 20% were prepared by an impregnation method. The as-synthesized Cu/MgO catalysts were characterized by N2 adsorption, XRD, TEM, CO2-TPD, XPS and TPR. These catalysts were found to be effective for ethanol dehydrogenation into acetaldehyde. As the Cu loading was increased, the ethanol conversion first increased and then leveled off. At a WHSV of 1.5 h−1 and 250 °C, the 20%Cu/MgO catalyst gave an initial conversion of 81.5%, with 97.7% selectivity toward acetaldehyde. Compared to 20%Cu/SiO2, the 20%Cu/MgO catalyst displayed an equivalent initial acetaldehyde yield, higher acetaldehyde selectivity and longer stability. Full article
(This article belongs to the Special Issue Catalytic Conversion of Renewable Biomass Platform Molecules)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of the catalysts: (<b>a</b>) 2.5%Cu/MgO, (<b>b</b>) 5%Cu/MgO, (<b>c</b>) 10%Cu/MgO, (<b>d</b>) 15%Cu/MgO and (<b>e</b>) 20%Cu/MgO and (<b>f</b>) 20%Cu/SiO<sub>2</sub>.</p>
Full article ">Figure 2
<p>(<b>a</b>) TEM image and (<b>b</b>) HAADF STEM mapping images of the 20%Cu/MgO catalyst.</p>
Full article ">Figure 3
<p>(<b>a</b>) TEM image and (<b>b</b>) HAADF STEM mapping images of the 20%Cu/SiO<sub>2</sub> catalyst.</p>
Full article ">Figure 4
<p>CO<sub>2</sub>-TPD profiles of the MgO support and Cu/MgO catalysts: (<b>a</b>) MgO, (<b>b</b>) 2.5%Cu/MgO, (<b>c</b>) 5%Cu/MgO, (<b>d</b>) 10%Cu/MgO, (<b>e</b>) 15%Cu/MgO and (<b>f</b>) 20%Cu/MgO.</p>
Full article ">Figure 5
<p>Cu 2p XPS spectra of the 20%CuO/MgO and 20%CuO/SiO<sub>2</sub> precursors as well as the reduced 20%Cu/MgO and 20%Cu/SiO<sub>2</sub> catalysts.</p>
Full article ">Figure 6
<p>Cu LMM XAES spectra of the 20%Cu/MgO and 20%Cu/SiO<sub>2</sub> catalysts.</p>
Full article ">Figure 7
<p>H<sub>2</sub>-TPR profiles of the CuO/MgO and CuO/SiO<sub>2</sub> precursors: (<b>a</b>) 2.5%CuO/MgO, (<b>b</b>) 5%CuO/MgO, (<b>c</b>) 10%CuO/MgO, (<b>d</b>) 15%CuO/MgO, (<b>e</b>) 20%CuO/MgO and (<b>f</b>) 20%CuO/SiO<sub>2</sub>.</p>
Full article ">Figure 8
<p>(<b>a</b>) Effect of the reaction temperature on the ethanol conversion and (<b>b</b>) selectivity to acetaldehyde over Cu/MgO catalysts with different Cu loadings (2.5–20%). Reaction condition: WHSV = 1.5 h<sup>−1</sup>, reaction from 190 to 290 °C and holding at each temperature for 1 h. Data were obtained after 1 h of the reaction.</p>
Full article ">Figure 9
<p>Relationship between the ethanol conversion obtained at 230 °C and the number of surface Cu atoms on the Cu/MgO catalysts. Data were obtained after 1 h of the reaction.</p>
Full article ">Figure 10
<p>Relationship between the productivity obtained at 230 °C and the dispersion of Cu particles on the MgO support. Data were obtained after 1 h of the reaction.</p>
Full article ">Figure 11
<p>Reuse of the 2.5%Cu/MgO catalyst for 2 runs. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H<sub>2</sub>/Ar. Reaction condition: 270 °C, WHSV = 1.5 h<sup>−1</sup>.</p>
Full article ">Figure 12
<p>Effect of the WHSV on the ethanol conversion and acetaldehyde selectivity over 20%Cu/MgO. Reaction condition: 250 °C, after 1 h of the reaction at each WHSV.</p>
Full article ">Figure 13
<p>Catalytic performance of the (<b>a</b>) 20%Cu/MgO and (<b>b</b>) 20%Cu/SiO<sub>2</sub> catalysts over a prolonged period of time for 2 runs’ reuse. Regeneration condition: oxidation at 500 °C for 1 h under an air stream followed by reduction at 300 °C for 1 h under a flow of 10 vol% H<sub>2</sub>/Ar. Reaction condition: 250 °C, WHSV = 1.5 h<sup>−1</sup>.</p>
Full article ">Figure 14
<p>XRD patterns of the fresh and used 20%Cu/MgO and 20%Cu/SiO<sub>2</sub> catalysts after reuse for 2 runs, as depicted in <a href="#catalysts-14-00541-f013" class="html-fig">Figure 13</a>.</p>
Full article ">
20 pages, 5431 KiB  
Article
Catalytic Ozonation of Sulfachloropyridazine Sodium by Diatomite-Modified Fe2O3: Mechanism and Pathway
by Yang Yu, Lingling Wang, Zhandong Wu, Xuguo Liu, Zhen Liu, Lijian Zhang and Lixin Li
Catalysts 2024, 14(8), 540; https://doi.org/10.3390/catal14080540 - 19 Aug 2024
Viewed by 462
Abstract
A diatomite-modified Fe2O3 (Fe2O3/Dia) catalyst was prepared to catalyze the ozonation degradation of sulfachloropyridazine sodium (SPDZ). The chemical oxygen demand (COD) was used as the index of pollutant degradation. The catalytic ozonation experiment showed that the [...] Read more.
A diatomite-modified Fe2O3 (Fe2O3/Dia) catalyst was prepared to catalyze the ozonation degradation of sulfachloropyridazine sodium (SPDZ). The chemical oxygen demand (COD) was used as the index of pollutant degradation. The catalytic ozonation experiment showed that the COD removal rate of SPDZ was 87% under Fe2O3/Dia catalysis, which was much higher than that obtained when using Fe2O3 as the catalyst. The characteristics of the Fe2O3/Dia catalyst were investigated, and the successful synthesis of the Fe2O3/Dia composite catalyst was proved by XRD, XPS, SEM, FTIR, BET and other characterization methods. The catalytic mechanism of degradation by ozone with Fe2O3/Dia was analyzed. According to free-radical trapping experiments and an in situ electron paramagnetic spectrometer characterization analysis, the main oxidizing species in the catalytic Fe2O3/Dia ozone system is ·OH. The intermediates in the degradation process of SPDZ were detected and analyzed in detail by liquid chromatography-coupled mass spectrometry. The degradation mechanism and three degradation paths of SPDZ were proposed. Full article
(This article belongs to the Section Environmental Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Effect of (<b>a</b>) catalyst dosage, (<b>b</b>) ozone dosages, (<b>c</b>) pH and (<b>d</b>) temperature on COD degradation of sulfonamide chlordazine sodium catalyzed by Fe<sub>2</sub>O<sub>3</sub>/Dia.</p>
Full article ">Figure 2
<p>COD degradation curve for various systems. Reaction conditions: [SPDZ]<sub>0</sub> = 2 g·L<sup>−1</sup>, catalyst dosage = 2 g·L<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Effect of (<b>a</b>) TBA, (<b>b</b>) HCO<sub>3</sub><sup>−</sup>, and (<b>c</b>) PBQ on the catalytic ozonation of SPDZ. Reaction conditions: [SPDZ]<sub>0</sub> = 2 g·L<sup>−1</sup>, catalyst dosage = 2 g·L<sup>−1</sup>.</p>
Full article ">Figure 4
<p>EPR signal of reactive oxygen species. (<b>a</b>) DMPO-·OH, (<b>b</b>) DMPO- O<sub>2</sub>·<sup>−</sup>.</p>
Full article ">Figure 4 Cont.
<p>EPR signal of reactive oxygen species. (<b>a</b>) DMPO-·OH, (<b>b</b>) DMPO- O<sub>2</sub>·<sup>−</sup>.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effect of PO<sub>4</sub><sup>3−</sup> on the catalytic ozonation of SPDZ with Fe<sub>2</sub>O<sub>3</sub>/Dia. (<b>b</b>) Effect of various PO<sub>4</sub><sup>3−</sup> concentrations on the catalytic ozonation of SPDZ with Fe<sub>2</sub>O<sub>3</sub>/Dia. Reaction conditions: [SPDZ]<sub>0</sub> = 2 g·L<sup>−1</sup>, catalyst dosage = 2 g·L<sup>−1</sup>.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>b</b>) SEM of diatomite before modification. (<b>c</b>,<b>d</b>) SEM of diatomite after modification. (<b>e</b>,<b>f</b>) SEM of diatomite-modified Fe<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 7
<p>XRD spectra of (<b>a</b>) diatomaceous earth, (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>, and (<b>c</b>) Fe<sub>2</sub>O<sub>3</sub>/Dia.</p>
Full article ">Figure 8
<p>XPS spectra of Fe<sub>2</sub>O<sub>3</sub>/Dia: (<b>a</b>) survey spectrum, (<b>b</b>) Fe 2p spectrum, (<b>c</b>) Si 2p spectrum.</p>
Full article ">Figure 9
<p>Mechanism diagram of ROS formation from ozone catalyzed by the Fe<sub>2</sub>O<sub>3</sub>/Dia composite catalyst.</p>
Full article ">Figure 10
<p>FTIR spectra of Fe<sub>2</sub>O<sub>3</sub>/Dia: (<b>a</b>) full spectrum, (<b>b</b>) selected range.</p>
Full article ">Figure 11
<p>The catalytic ozonation pathway of SPDZ was proposed.</p>
Full article ">Figure 12
<p>Recycling tests of SPDZ degradation by Fe<sub>2</sub>O<sub>3</sub>/Dia-catalyzed ozonation.</p>
Full article ">
14 pages, 5631 KiB  
Article
Strengthened Removal of Tetracycline by a Bi/Ni Co-Doped SrTiO3/TiO2 Composite under Visible Light
by Weifang Chen, Na Zhao, Mingzhu Hu, Xingguo Liu and Baoqing Deng
Catalysts 2024, 14(8), 539; https://doi.org/10.3390/catal14080539 - 19 Aug 2024
Viewed by 410
Abstract
A two-step hydrothermal method was used to first obtain a SrTiO3/TiO2 composite then to dope the composite with Bi, Ni and Bi/Ni. Morphology, crystalline structures, surface valances and optical features of SrTiO3/TiO2 and Bi-, Ni-, Bi/Ni-doped SrTiO [...] Read more.
A two-step hydrothermal method was used to first obtain a SrTiO3/TiO2 composite then to dope the composite with Bi, Ni and Bi/Ni. Morphology, crystalline structures, surface valances and optical features of SrTiO3/TiO2 and Bi-, Ni-, Bi/Ni-doped SrTiO3/TiO2 were assessed. XRD and XPS analysis showed that Bi and Ni were successfully doped and existed in Bi(3+) and Ni(2+) oxidation state. UV–vis analysis further revealed that the bandgap energies of TiO2 and SrTiO3/TiO2 were calculated to be 3.14 eV and 3.04 eV. By comparison, Bi, Ni and Bi/Ni doping resulted in the narrowing of bandgaps to 2.82 eV, 2.96 eV and 2.69 eV, respectively. The removal ability of SrTiO3/TiO2 and doped SrTiO3/TiO2 were investigated with tetracycline as the representative pollutant. After 40 min of exposure to visible light, Bi/Ni co-doped SrTiO3/TiO2 photocatalyst was able to remove 90% of the tetracycline with a mineralization rate of about 70%. In addition, first-order removal rate constant was 0.0074 min−1 for SrTiO3/TiO2 and increased to 0.0278 min−1 after co-doping. The strengthened removal by co-doped photocatalyst was attributed mainly to the enhanced absorption of visible light as co-doping resulted in the decreases of bandgap energies. At the same time, the co-doped material was robust against changes in pH. Removal of tetracycline was stable as pH changed from 5 to 9. Tetracycline removal was inhibited to a certain degree by the presence of nitrate, phosphate and high concentration of humic acid. Moreover, the co-doped material exhibited strong structural stability and reusability. In addition, a photocatalysis mechanism with photogenerated holes and ·O2 radicals as main oxidative species was proposed based on entrapping experiments and EPR results. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the photocatalysts: (<b>a</b>) TiO<sub>2</sub>; (<b>b</b>)S-TO; (<b>c</b>) Bi/S-TO; (<b>d</b>) Ni/S-TO; (<b>e</b>) Bi/Ni/S-TO.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern (<b>b</b>) Locally enlarged XRD patterns of TiO<sub>2</sub>, S-TO, Bi/S-TO, Ni-S-TO and Bi/Ni/S-TO.</p>
Full article ">Figure 3
<p>XPS Spectra: (<b>a</b>) Bi/Ni/S-TO; (<b>b</b>) O 1s; (<b>c</b>) Ti 2p; (<b>d</b>) Sr 3d; (<b>e</b>) Bi 4f; (<b>f</b>) Ni 2p.</p>
Full article ">Figure 3 Cont.
<p>XPS Spectra: (<b>a</b>) Bi/Ni/S-TO; (<b>b</b>) O 1s; (<b>c</b>) Ti 2p; (<b>d</b>) Sr 3d; (<b>e</b>) Bi 4f; (<b>f</b>) Ni 2p.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–vis absorption spectra of TiO<sub>2</sub>, S-TO, Bi/S-TO, Ni-S-TO and Bi/Ni/S-TO and (<b>b</b>) bandgap energies of TiO<sub>2</sub> and Bi/Ni/S-TO.</p>
Full article ">Figure 5
<p>Tetracycline removal: (<b>a</b>) Comparison of photocatalyst; (<b>b</b>) Effects of dosage; (<b>c</b>) Effects of initial pH; (<b>d</b>) Effects of co-existing anion; (<b>e</b>) Effects of co-existing humic acid; (<b>f</b>) Mineralization; (<b>g</b>) Stability and reusability; (<b>h</b>) XRD spectra.</p>
Full article ">Figure 6
<p>(<b>a</b>) Free radical trapping (<b>b</b>) EPR signal of DMPO-·O<sub>2</sub><sup>−</sup> and (<b>c</b>) EPR signal of TEMPO-h<sup>+</sup>.</p>
Full article ">Figure 7
<p>(<b>a</b>) VB-XPS spectrum of Bi/Ni/S-TO. (<b>b</b>) Possible reaction mechanism of tetracycline removal by Bi/Ni/S-TO.</p>
Full article ">
25 pages, 3741 KiB  
Review
Exploring Intermetallic Compounds: Properties and Applications in Catalysis
by Zhiquan Hou, Mengwei Hua, Yuxi Liu, Jiguang Deng, Xin Zhou, Ying Feng, Yifan Li and Hongxing Dai
Catalysts 2024, 14(8), 538; https://doi.org/10.3390/catal14080538 - 18 Aug 2024
Viewed by 814
Abstract
Intermetallic compounds (IMCs) have attracted significant attention in recent years due to their unique properties and potential applications in various fields, particularly in catalysis. This review aims to provide an in-depth understanding of IMCs, including their synthesis methods, structural characteristics, and diverse catalytic [...] Read more.
Intermetallic compounds (IMCs) have attracted significant attention in recent years due to their unique properties and potential applications in various fields, particularly in catalysis. This review aims to provide an in-depth understanding of IMCs, including their synthesis methods, structural characteristics, and diverse catalytic applications. The review begins with an introduction to IMCs, highlighting their distinct features and advantages over traditional catalyst materials. It then delves into the synthesis techniques employed to prepare IMCs and explores their structural properties. Subsequently, catalytic applications of the IMCs are introduced, focusing on the key reactions and highlighting their superior catalytic performance compared to conventional catalysts. Future perspectives for, and challenges to, the catalysis of IMCs are then proposed. Full article
Show Figures

Figure 1

Figure 1
<p>Preparation strategies and applications of IMCs in catalysis.</p>
Full article ">Figure 2
<p>Schematic illustration of bimetallic alloys structure [<a href="#B11-catalysts-14-00538" class="html-bibr">11</a>]. Reprinted with permission from Ref. [<a href="#B11-catalysts-14-00538" class="html-bibr">11</a>]. Copyright 2017, copyright American Chemical Society.</p>
Full article ">Figure 3
<p>The ORR activity volcano plot of four ordered Pt<sub>2</sub>CoM, ordered PtCo, and five disordered PtCo. [<a href="#B24-catalysts-14-00538" class="html-bibr">24</a>] Reprinted with permission from Ref. [<a href="#B24-catalysts-14-00538" class="html-bibr">24</a>]. Copyright 2024, copyright Springer Nature.</p>
Full article ">Figure 4
<p>Schematic illustration of the arc-melting synthesis of the structurally ordered Pt–M (M = Fe, Co, and Ni) IMCs (IMCs) [<a href="#B46-catalysts-14-00538" class="html-bibr">46</a>]. Reprinted with permission from Ref. [<a href="#B46-catalysts-14-00538" class="html-bibr">46</a>]. Copyright 2022, copyright Springer Nature.</p>
Full article ">Figure 5
<p>Schematic illustration of different syntheses of intermetallic compounds.</p>
Full article ">Figure 6
<p>Schematic illustrations of traditional synthetic strategies for ordered intermetallic nanocrystals [<a href="#B27-catalysts-14-00538" class="html-bibr">27</a>]. Reprinted with permission from Ref. [<a href="#B27-catalysts-14-00538" class="html-bibr">27</a>]. Copyright 2023, copyright Springer Nature.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic preparation process of the PdZn–sub-2@ZIF-8C using a MOF-confined co-reduction strategy, (<b>b</b>) HAADF–STEM image of PdZn–1.2@ZIF-8C, (<b>c</b>) HAADF–STEM image of PdZn–1.8@ZIF-8C, (<b>d</b>) HAADF–STEM image of PdZn–2.7/ZIF-8C, and the insets of panels (<b>b</b>–<b>d</b>) are the corresponding histograms of particle-size distributions of the PdZn–1.2@ZIF-8C, PdZn–1.8@ZIF-8C, and PdZn–2.7/ZIF-8C, respectively, (<b>e</b>) EDS elementary mapping images of the PdZn–1.2@ZIF-8C, scale bar = 20 nm, and (<b>f</b>) high-resolution HAADF–STEM image of PdZn–10/ZIF-8C [<a href="#B69-catalysts-14-00538" class="html-bibr">69</a>]. Reprinted with permission from Ref. [<a href="#B69-catalysts-14-00538" class="html-bibr">69</a>]. Copyright 2018, copyright John Wiley and Sons.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic illustration of gas-phase ordered alloying strategy for the preparation of Pt<sub>1</sub>Fe<sub>1</sub>/Fe<sub>1</sub>–N–C and conventional thermal annealing strategy for the preparation of contrast sample PtFe&amp;Fe/Fe<sub>1</sub>–N–C, (<b>b</b>) TEM image of pre-synthesized Pt nanoparticles/NC, (<b>c</b>) TEM, (<b>d</b>) HRTEM images of Pt<sub>1</sub>Fe<sub>1</sub>/Fe<sub>1</sub>–N–C, (<b>e</b>,<b>f</b>) atomic-resolution HAADF–STEM image and corresponding FFT pattern of Pt<sub>1</sub>Fe<sub>1</sub> IMC, (<b>g</b>) schematic diagram of the Pt<sub>1</sub>Fe<sub>1</sub> IMC structure, (<b>h</b>) aberration-corrected HAADF–STEM image, and (<b>i</b>) EDS elemental mappings of Pt<sub>1</sub>Fe<sub>1</sub>/Fe<sub>1</sub>–N–C [<a href="#B72-catalysts-14-00538" class="html-bibr">72</a>]. Reprinted with permission from Ref. [<a href="#B72-catalysts-14-00538" class="html-bibr">72</a>]. Copyright 2023, copyright John Wiley and Sons.</p>
Full article ">Figure 9
<p>Catalytic activity of NaAu<sub>2</sub> for CO oxidation. Reaction rates were measured in a plug-flow reactor with 150 mg of NaAu<sub>2</sub> powders. The reaction gases were composed of CO (14 mL/min), O<sub>2</sub> (35 mL/min), and He (31 mL/min) at 1 atm, and the Arrhenius plot showed that the activation energy of NaAu<sub>2</sub> for CO oxidation was 31.6 ± 0.7 kJ/mol (inset) [<a href="#B83-catalysts-14-00538" class="html-bibr">83</a>]. Reprinted with permission from Ref. [<a href="#B83-catalysts-14-00538" class="html-bibr">83</a>]. Copyright 2013, copyright American Chemical Society.</p>
Full article ">
10 pages, 2068 KiB  
Article
Catalytic Effect of Alkali Metal Ions on the Generation of CO and CO2 during Lignin Pyrolysis: A Theoretical Study
by Xiaoyan Jiang, Yiming Han, Baojiang Li, Ji Liu, Guanzheng Zhou, Xiaojiao Du, Shougang Wei, Hanxian Meng and Bin Hu
Catalysts 2024, 14(8), 537; https://doi.org/10.3390/catal14080537 - 18 Aug 2024
Viewed by 489
Abstract
A density functional theory method was employed to conduct theoretical calculations on the pyrolysis reaction pathways of lignin monomer model compounds with an aldehyde or carboxyl group under the catalytic effect of alkali metal ions Na+ and K+, exploring their [...] Read more.
A density functional theory method was employed to conduct theoretical calculations on the pyrolysis reaction pathways of lignin monomer model compounds with an aldehyde or carboxyl group under the catalytic effect of alkali metal ions Na+ and K+, exploring their influence on the formation of the small molecular gaseous products CO and CO2. The results indicate that Na+ and K+ can easily bind with the oxygen-containing functional groups of the lignin monomer model compounds to form stable and low-energy complexes. Except for benzaldehyde and p-hydroxybenzaldehyde, Na+ and K+ can facilitate the decarbonylation reactions of other benzaldehyde-based and phenylacetaldehyde-based lignin monomer model compounds during the pyrolysis process, thereby enhancing the generation of CO. When the characteristic functional groups on the benzene rings of benzaldehyde-based and phenylacetaldehyde-based lignin monomer model compounds are the same, the phenylacetaldehyde-based ones are more prone to undergo decarbonylation than the benzaldehyde-based ones. Additionally, both Na+ and K+ can inhibit the decarboxylation reactions of benzoic acid-based and phenylacetic acid-based lignin monomer model compounds, thereby restraining the formation of CO2. When the characteristic functional groups on the benzene rings of benzoic acid-based and phenylacetic acid-based lignin monomer model compounds are the same, the phenylacetic acid-based ones are more difficult to undergo decarboxylation than the benzoic acid-based ones. Full article
(This article belongs to the Collection Catalytic Conversion of Biomass to Bioenergy)
Show Figures

Figure 1

Figure 1
<p>Lignin monomer model compounds containing aldehyde and carboxyl groups at the C<sub>α</sub> and C<sub>β</sub> sites.</p>
Full article ">Figure 2
<p>The optimal geometric configurations of the lignin monomer model compounds M<sub>1x</sub> and M<sub>2x</sub> catalyzed by alkali metal ions. (unit: nm).</p>
Full article ">Figure 2 Cont.
<p>The optimal geometric configurations of the lignin monomer model compounds M<sub>1x</sub> and M<sub>2x</sub> catalyzed by alkali metal ions. (unit: nm).</p>
Full article ">Figure 3
<p>The optimal geometric configurations of the lignin monomer model compounds M<sub>3x</sub> and M<sub>4x</sub> catalyzed by alkali metal ions. (unit: nm).</p>
Full article ">Figure 4
<p>Decarbonylation reaction pathways of lignin monomer model compounds M<sub>1x</sub> and M<sub>2x</sub>.</p>
Full article ">Figure 5
<p>Decarboxylation reaction pathways of lignin monomer model compounds M<sub>3x</sub> and M<sub>4x</sub>.</p>
Full article ">
18 pages, 4714 KiB  
Article
α-L-Arabinofuranosidases of Glycoside Hydrolase Families 43, 51 and 62: Differences in Enzyme Substrate and Positional Specificity between and within the GH Families
by Walid Fathallah and Vladimír Puchart
Catalysts 2024, 14(8), 536; https://doi.org/10.3390/catal14080536 - 17 Aug 2024
Viewed by 423
Abstract
The increasing number of uncharacterized proteins in the CAZy database highlights the importance of their functional characterization. Therefore, the substrate and positional specificity of 34 α-L-arabinofuranosidases classified into GH43, GH51, and GH62 families was determined on arabinoxylan, arabinan, and derived oligosaccharides (many enzyme–substrate [...] Read more.
The increasing number of uncharacterized proteins in the CAZy database highlights the importance of their functional characterization. Therefore, the substrate and positional specificity of 34 α-L-arabinofuranosidases classified into GH43, GH51, and GH62 families was determined on arabinoxylan, arabinan, and derived oligosaccharides (many enzyme–substrate combinations were examined for the first time) covering all possible kinds of arabinofuranosyl branches using TLC. Arabinoxylan was efficiently debranched by the majority of the tested proteins. Most of them showed AXH-m specificity, acting on 2- or 3-monoarabinosylated substrates, while AXH-d3 specificity (liberation of 3-linked arabinose solely from 2,3 doubly decorated substrates) was found mainly in the subfamily GH43_10, harbouring enzymes of both types. Several GH51 enzymes, however, released arabinose also from a xylooligosaccharide doubly arabinosylated at the non-reducing end. The AXH-m and AXH-d3 specificities correlated well with the dearabinosylation of arabinan and arabinooligosaccharides, which were debranched by all GH51 representatives and some GH43 and GH62 members. The GH51 and GH62 arabinan-debranching enzymes also hydrolyzed debranched arabinan, while within the GH43 family the linear arabinan-degrading ability was found only in the GH43_26 subfamily, comprising specific exo 1,5-α-L-arabinofuranosidases. This study demonstrates a first attempt in the systematic examination of a relationship between CAZy classification and substrate and positional specificities of various α-L-arabinofuranosidases. Full article
(This article belongs to the Section Biocatalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>TLC analysis of wheat arabinoxylan (WAX) hydrolysates. A schematic structure of the polysaccharide in non-abbreviated and abbreviated forms is shown above the TLC plate. The plate was developed in the solvent system of 1-butanol/ethanol/water (10:8:5, by vol.; system 1), and the sugars were visualized using orcinol detection reagent. The GH43, GH51, and GH62 enzymes are marked in black, blue, and green colours, respectively. Aliquots of the reaction mixtures were spotted after 1 and/or 4 days of incubation.</p>
Full article ">Figure 2
<p>TLC analysis of hydrolysates of branched arabinoxylooligosaccharide mixture comprising 3<sup>3</sup>-α-L-arabinofuranosyl-xylotetraose (XA<sup>3</sup>XX) and 2<sup>3</sup>-α-L-arabinofuranosyl-xylotetraose (XA<sup>2</sup>XX). The plate was developed twice in the solvent system of 1-propanol/ethanol/water (7:1:2, by vol.; system 3), and the sugars were visualized using orcinol detection reagent. The GH43, GH51, and GH62 enzymes are marked in black, blue, and green colours, respectively. Aliquots of the reaction mixtures were spotted after 4 h, 1, and/or 4 days of incubation.</p>
Full article ">Figure 3
<p>Comparison of amino acid sequences of the proteins used in this study classified into GH43 family. Enzymes belonging to subfamilies that are represented by multiple members used in this work were only chosen for this comparison. Panel (<b>A</b>): multiple sequence alignment of the sequences. The enzymes with solved 3-D structures are marked with a <span>$</span> symbol (accessible under PDB codes 5A8C and 3QEF for <span class="html-italic">Bt</span>Abf43B and <span class="html-italic">Cj</span>Abf43A). The asterisk at the end of the <span class="html-italic">Bt</span>Abf43D sequence denotes C-terminus of the protein. Catalytically competent amino acids, i.e., catalytic base, catalytic acid, and modulator of catalytic acid, are marked in violet, green, and cyan, respectively. The residues forming subsites −1 and +1 are highlighted in yellow and red, respectively. The residues interacting with the substrate in other subsites are indicated in grey. Panel (<b>B</b>): phylogenetic tree corresponding to the multiple sequence alignment. Classification of the enzymes into the GH43 subfamilies is indicated.</p>
Full article ">Figure 4
<p>TLC analysis of sugar beet arabinan (SBA) hydrolysates. Schematic structure of the polysaccharide in non-abbreviated and abbreviated forms is shown above the TLC plate. The plate was developed in the solvent system of 1-butanol/ethanol/water (10:8:5, by vol.; system 1), and the sugars were visualized using orcinol detection reagent. The GH43, GH51, and GH62 enzymes are marked in black, blue, and green colours, respectively. Aliquots of the reaction mixtures were spotted after 1 and/or 4 days of incubation.</p>
Full article ">Figure 5
<p>TLC analysis of hydrolysates of branched arabinooligosaccharide mixture comprising 2<sup>2</sup>,3<sup>2</sup>-di-α-L-arabinofuranosyl-(1,5)-α-L-arabinotriose (AA<sup>2,3</sup>A) and 3<sup>2</sup>-α-L-arabinofuranosyl-(1,5)-α-L-arabinotetraose (AAA<sup>3</sup>A). The plate was developed twice in the solvent system of chloroform/acetic acid/water (6:7:1, by vol.; system 2), and the sugars were visualized using orcinol detection reagent. The GH43, GH51. and GH62 enzymes are marked in black, blue. and green colours, respectively. Aliquots of the reaction mixtures were spotted after 4 h, 1, and/or 4 days of incubation.</p>
Full article ">Figure 6
<p>TLC analysis of debranched arabinan (DA) hydrolysates. The schematic structure of the polysaccharide is shown above the TLC plate. The plate was developed in the solvent system of 1-butanol/ethanol/water (10:8:5, by vol.; system 1), and the sugars were visualized using orcinol detection reagent. The GH43, GH51, and GH62 enzymes are marked in black, blue, and green colours, respectively. Aliquots of the reaction mixtures were spotted after 2 h, 1, and/or 4 days of incubation.</p>
Full article ">
15 pages, 2562 KiB  
Article
Sludge Recycling from Non-Lime Purification of Electrolysis Wastewater: Bridge from Contaminant Removal to Waste-Derived NOX SCR Catalyst
by Ju Gao, Fucheng Sun, Pei Liu, Jizhi Zhou and Yufeng Zhang
Catalysts 2024, 14(8), 535; https://doi.org/10.3390/catal14080535 - 17 Aug 2024
Viewed by 661
Abstract
Catalysts for the selective catalytic reduction (NOX SCR) of nitrogen oxides can be obtained from sludge in industrial waste treatment, and, due to the complex composition of sludge, NOX SCR shows various SCR efficiencies. In the current work, an SCR catalyst [...] Read more.
Catalysts for the selective catalytic reduction (NOX SCR) of nitrogen oxides can be obtained from sludge in industrial waste treatment, and, due to the complex composition of sludge, NOX SCR shows various SCR efficiencies. In the current work, an SCR catalyst developed from the sludge produced with Fe/C micro-electrolysis Fenton technology (MEF) in wastewater treatment was investigated, taking into account various sludge compositions, Fe/C ratios, and contaminant contents. It was found that, at about 300 °C, the NOX removal rate could reach 100% and there was a wide decomposition temperature zone. The effect of individual components of electroplating sludge, i.e., P, Fe and Ni, on NOX degradation performance of the obtained solids was investigated. It was found that the best effect was achieved when the Fe/P was 8/3 wt%, and variations in the Ni content had a limited effect on the NOX degradation performance. When the Fe/C was 1:2 and the Fe/C/P was 1:2:0.4, the electroplating sludge formed after treatment with Fe/C MEF provided the best NOX removal rate at 100%. Moreover, the characterization results show that the activated carbon was also involved in the catalytic reduction degradation of NOX. An excessive Fe content may cause agglomeration on the catalyst surface and thus affect the catalytic efficiency. The addition of P effectively reduces the catalytic reaction temperature, and the formation of phosphate promotes the generation of adsorbed oxygen, which in turn contributes to improvements in catalytic efficiency. Therefore, our work suggests that controlling the composition in the sludge is an efficient way to modulate SCR catalysis, providing a bridge from contaminant-bearing waste to efficient catalyst. Full article
(This article belongs to the Special Issue Homogeneous and Heterogeneous Catalytic Oxidation and Reduction)
Show Figures

Figure 1

Figure 1
<p>Fe/C + H<sub>2</sub>O<sub>2</sub> treatment of electroplating wastewater effluent in (<b>a</b>) TP, Ni, and Fe<sup>2+</sup> and (<b>b</b>) pH change.</p>
Full article ">Figure 2
<p>The effect of composition on NO<sub>X</sub> removal efficiency.</p>
Full article ">Figure 3
<p>The effects of (<b>a</b>) P, (<b>b</b>) Ni, and (<b>c</b>) Fe on NO<sub>X</sub> removal efficiency.</p>
Full article ">Figure 4
<p>The effects of (<b>a</b>) Fe/C and (<b>b</b>) P concentration in the pollutants on NO<sub>X</sub> removal efficiency.</p>
Full article ">Figure 5
<p>SEM images of (<b>a</b>) Fe<sub>8</sub>/P<sub>4</sub>/Ni/C (10.0 μm); (<b>b</b>) Fe<sub>8</sub>/P<sub>4</sub>/Ni/C (1.0 μm); (<b>c</b>) Fe<sub>10</sub>/P<sub>5</sub>/Ni/C (10.0 μm); and (<b>d</b>) Fe<sub>10</sub>/P<sub>5</sub>/Ni/C samples (1.0 μm).</p>
Full article ">Figure 6
<p>(<b>a</b>) The NH<sub>3</sub> TPD curves of Fe, Fe/Ni and Fe/Ni/P/C; (<b>b</b>) the NH<sub>3</sub> TPD curves of Fe/Ni/P/C samples with different P contents; and (<b>c</b>) the NO<sub>2</sub> TPD curves of the sample before and after reaction.</p>
Full article ">Figure 7
<p>Schematic of the reaction.</p>
Full article ">Figure 8
<p>The NO<sub>X</sub> catalytic experimental device and technological process.</p>
Full article ">
12 pages, 3975 KiB  
Article
Facile Synthesis of Ni(OH)2 through Low-Temperature N-Doping for Efficient Hydrogen Evolution
by Zi-Zhang Liu, Ruo-Yao Fan, Ning Yu, Ya-Nan Zhou, Xin-Yu Zhang, Bin Dong and Zi-Feng Yan
Catalysts 2024, 14(8), 534; https://doi.org/10.3390/catal14080534 - 16 Aug 2024
Viewed by 390
Abstract
Nickel hydroxide is a potentially cheap non-precious metal catalytic material for alkaline hydrogen evolution reactions (HERs). Herein, a nickel form (NF)-based nitrogen-modified nickel hydroxide (N-Ni(OH)2/NF) with interlaced two-dimensional (2D) nanosheet structures was synthesized by a simple one-step ammonia vapor-phase hydrothermal method [...] Read more.
Nickel hydroxide is a potentially cheap non-precious metal catalytic material for alkaline hydrogen evolution reactions (HERs). Herein, a nickel form (NF)-based nitrogen-modified nickel hydroxide (N-Ni(OH)2/NF) with interlaced two-dimensional (2D) nanosheet structures was synthesized by a simple one-step ammonia vapor-phase hydrothermal method for efficient electrocatalytic HERs. The effect of the reaction temperature of the catalyst preparation on the HERs’ performance was studied in detail. The HER activity of N-Ni(OH)2/NF is enhanced by the large specific surface area, mass transfer and electron conductivity provided by a unique and suitable 2D nanostructure and nitrogen doping. The obtained N-Ni(OH)2/NF not only shows a superior HERs performance, but also exhibits good stability during long-term electrolysis. Full article
(This article belongs to the Special Issue Non-novel Metal Electrocatalytic Materials for Clean Energy)
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of one-step access to N-Ni(OH)<sub>2</sub>/NF.</p>
Full article ">Figure 2
<p>XRD patterns of N<sub>90</sub>-Ni(OH)<sub>2</sub>/NF, N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF and N<sub>180</sub>-Ni(OH)<sub>2</sub>/NF.</p>
Full article ">Figure 3
<p>XPS spectra of N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF: (<b>a</b>) survey; (<b>b</b>) N 1s; (<b>c</b>) Ni 2p.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>,<b>b</b>) N<sub>90</sub>-Ni(OH)<sub>2</sub>/NF; (<b>c</b>,<b>d</b>) N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF and (<b>e</b>,<b>f</b>) N<sub>180</sub>-Ni(OH)<sub>2</sub>/NF.</p>
Full article ">Figure 5
<p>(<b>a</b>) TEM and (<b>b</b>) HRTEM images of N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF and (<b>c</b>) corresponding element mapping images.</p>
Full article ">Figure 6
<p>Electrocatalytic measurements for HER at 1.0 M KOH. (<b>a</b>) Linear sweep voltammogram (LSV). (<b>b</b>) Tafel plots. (<b>c</b>) Electrochemical impedance spectroscopy (EIS). (<b>d</b>) Determined double-layer capacitance (C<sub>dl</sub>).</p>
Full article ">Figure 7
<p>The Faraday efficiency for HER on N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF in 1 M KOH.</p>
Full article ">Figure 8
<p>Stability test of N<sub>130</sub>-Ni(OH)<sub>2</sub>/NF. (<b>a</b>) LSV before and after cyclic voltammograms (CVs) for 5000 cycles. (<b>b</b>) 12 h CA test.</p>
Full article ">
18 pages, 4266 KiB  
Article
Selective Oligomerization of Isobutylene in Mixed C4 with Co/BETA-Loaded Molecular Sieve Catalysts
by Xiaoping Chen, Panhu Yu, Hui Tian and Shuguang Xiang
Catalysts 2024, 14(8), 533; https://doi.org/10.3390/catal14080533 - 16 Aug 2024
Viewed by 366
Abstract
This paper investigates the use of loaded Co/BETA molecular sieve catalysts for the selective oligomerization of isobutylene. The physicochemical properties of Co/BETA molecular sieves were characterized using XRD, BET, NH3-TPD, FT-IR, XPS, and Py-FTIR. The effects of different active component loadings, reaction temperatures, [...] Read more.
This paper investigates the use of loaded Co/BETA molecular sieve catalysts for the selective oligomerization of isobutylene. The physicochemical properties of Co/BETA molecular sieves were characterized using XRD, BET, NH3-TPD, FT-IR, XPS, and Py-FTIR. The effects of different active component loadings, reaction temperatures, and reaction air velocities on the selective oligomerization of isobutylene were investigated in a fixed-bed reactor. The results showed that the catalytic effect was optimal when the Co loading was 6%, the reaction temperature was 60 °C, the reaction pressure was 1 MPa, and the reaction air speed was 1 h−1. The isobutylene conversion was greater than 74%, the C8= selectivity was approximately 70%, and the C8= yield reached 51.69% with minimal loss of n-butene, providing good catalytic capacity and efficiency. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of oligomerization reaction.</p>
Full article ">Figure 2
<p>XRD spectra of catalysts with different active component loadings.</p>
Full article ">Figure 3
<p>Adsorption–desorption isotherms (<b>A</b>) and pore size distribution maps (<b>B</b>) of catalysts with different Co loading levels.</p>
Full article ">Figure 4
<p>NH<sub>3</sub>-TPD plots for catalysts with different active components.</p>
Full article ">Figure 5
<p>FT-IR plots of catalysts with different active components.</p>
Full article ">Figure 6
<p>Py-FTIR spectra of (<b>A</b>) 0% Co and (<b>B</b>) 6% Co catalysts.</p>
Full article ">Figure 7
<p>Scanning spectra of 6% Co catalyst (<b>A</b>) and Co 2p region (<b>B</b>).</p>
Full article ">Figure 7 Cont.
<p>Scanning spectra of 6% Co catalyst (<b>A</b>) and Co 2p region (<b>B</b>).</p>
Full article ">Figure 8
<p>Evaluation results of loadings with different active components.</p>
Full article ">Figure 9
<p>Catalytic results at different reaction temperatures for catalysts with 6% loading of active components.</p>
Full article ">Figure 10
<p>Evaluation results of catalysts with 6% active component loading at different reaction air velocities.</p>
Full article ">Figure 11
<p>Fixed-bed reactor.</p>
Full article ">
18 pages, 10151 KiB  
Article
Application of Different Waveforms of Pulsed Current in the Classical Electro-Cocatalytic Process for Effective Removal of Sulfamethoxazole: Oxidation Mechanisms
by Jingkai Fang, Yongjian Wang, Jiahao Wang, Igor Ying Zhang and Rongfu Huang
Catalysts 2024, 14(8), 532; https://doi.org/10.3390/catal14080532 - 16 Aug 2024
Viewed by 548
Abstract
In this study, sulfamethoxazole (SMX) was applied as the model pollutant to assess the performance of pulsed current (PC) waveforms in the decontamination efficiency of the PC/peroxymonosulfate (PMS)/Fe(III) process and to investigate underlying oxidation mechanisms. Among the various waveforms tested, the sinusoidal wave [...] Read more.
In this study, sulfamethoxazole (SMX) was applied as the model pollutant to assess the performance of pulsed current (PC) waveforms in the decontamination efficiency of the PC/peroxymonosulfate (PMS)/Fe(III) process and to investigate underlying oxidation mechanisms. Among the various waveforms tested, the sinusoidal wave (SIN), combined with the Dimensionally Stable Anode (DSA) electrode, demonstrated superior degradation performance, with the order being SIN > ramp > square > direct current (DC). The operational conditions for the SIN/PMS/Fe(III) system were optimized to an initial pH of 3, a voltage of 6 V, 0.6 mmol/L of Fe3+, 1.0 mmol/L of PMS, and a frequency of 1 kHz. The results of quenching and probe experiments confirmed the generation of abundant reactive radicals such as OH, SO4•−, O2•−, Fe(IV), and 1O2 in the SIN/PMS/Fe(III) process, which collectively enhanced the degradation of SMX. Additionally, results of high-resolution mass spectrometry analysis were employed to identify the SMX oxidation byproducts, and the toxicity of SMX byproducts was evaluated. Overall, the SIN/PMS/Fe(III) process exhibits effective degradation capacity with high energy efficiency, establishing itself as an effective strategy for the practical treatment of medical wastewater. Full article
(This article belongs to the Section Industrial Catalysis)
Show Figures

Figure 1

Figure 1
<p>Comparison of performance of different waveform currents on SMX degradation regarding (<b>a</b>) removal of SMX and (<b>b</b>) calculated rate constants (the relationship between the molar concentration of reactants and the rate of a chemical reaction) in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na2SO4] = 20 mmol/L, voltage = 6 V, pulsed frequency = 1 kHz, duty cycle = 50%. Duty cycle refers to the percentage of time during the whole circle that the pulse is applied.</p>
Full article ">Figure 2
<p>Impact of different pH values (<b>a</b>) in SMX degradation using sinusoidal waveform currents with calculated rate constant k<sub>obs</sub> in (<b>d</b>). Impact of different voltages (<b>b</b>) in SMX degradation using sinusoidal waveform currents with calculated rate constant k<sub>obs</sub> in (<b>e</b>). Impact of different frequencies (<b>c</b>) in SMX degradation using sinusoidal waveform currents with calculated rate constant k<sub>obs</sub> in (<b>f</b>). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">Figure 3
<p>Impact of different concentrations of Fe(III) in (<b>a</b>,<b>c</b>) in SMX degradation using sinusoidal waveform currents with calculated rate constant k<sub>obs</sub> in (<b>b</b>,<b>d</b>). Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">Figure 4
<p>The concentrations of dissolved Fe(II) and total iron ions (<b>a</b>) in SIN/PMS/Fe(III) processes. The concentration of PMS (<b>b</b>) during the experiment in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">Figure 5
<p>Comparisons of the performance of different oxidation processes were conducted regarding the (<b>a</b>) removal of SMX and (<b>b</b>) calculated rate constants in different oxidation processes for SMX treatment. Reaction conditions: [SMX] = 2 mg/L, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L, voltage = 6 V, duty cycle = 50%, pulsed frequency = 1 kHz.</p>
Full article ">Figure 6
<p>EPR spectra for the detection of (<b>a</b>) <sup>•</sup>OH and SO<sub>4</sub><sup>•−</sup> and (<b>b</b>) <sup>1</sup>O<sub>2</sub> in the presence of DMPO and TEMP, respectively. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">Figure 7
<p>Quenching effects of different scavengers (<b>a</b>) on SMX degradation in SIN/PMS/Fe(III) system. Probe product 7-HC for <sup>•</sup>OH (<b>b</b>) in SIN/PMS/Fe(III) system. Probe product p-HBA for SO<sub>4</sub><sup>•−</sup> and PMSO<sub>2</sub> for Fe(IV) (<b>c</b>) in SIN/PMS/Fe(III) system. Probe product MF and DF for superoxide radical (<b>d</b>) in SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">Figure 8
<p>Proposed degradation pathways of SMX in SIN/PMS/Fe(III) system.</p>
Full article ">Figure 9
<p>Toxicity evaluation results of (<b>a</b>) acute toxicity LD<sub>50</sub>, (<b>b</b>) developmental toxicity, (<b>c</b>) mutagenicity, and (<b>d</b>) bioaccumulation factor of SMX and SMX byproducts in SIN/PMS/Fe(III) process obtained using the Toxicity Estimation Software Tool (T.E.S.T.).</p>
Full article ">Figure 10
<p>Influence of different anions including (<b>a</b>) Cl<sup>−</sup>, (<b>b</b>) NO<sub>3</sub><sup>−</sup>, (<b>c</b>) H<sub>2</sub>PO<sub>4</sub><sup>−</sup>, and (<b>d</b>) humic acid (HA) and (<b>e</b>) different water backgrounds for SMX degradation in the PE/PMS/Fe(III) system. (<b>f</b>) SMX degradation in five consecutive runs to examine the reusability of the SIN/PMS/Fe(III) system. Reaction conditions: [SMX] = 2 mg/L, voltage = 6 V, pulsed frequency = 1 kHz, [PMS] = 1.0 mmol/L, [Fe(III)] = 0.6 mmol/L, pH = 3, [Na<sub>2</sub>SO<sub>4</sub>] = 20 mmol/L.</p>
Full article ">
18 pages, 3108 KiB  
Review
The Advancement of Supported Bimetallic Catalysts for the Elimination of Chlorinated Volatile Organic Compounds
by Hongxia Lin, Yuxi Liu, Jiguang Deng, Lin Jing, Zhiwei Wang, Lu Wei, Zhen Wei, Zhiquan Hou, Jinxiong Tao and Hongxing Dai
Catalysts 2024, 14(8), 531; https://doi.org/10.3390/catal14080531 - 16 Aug 2024
Viewed by 496
Abstract
Chlorinated volatile organic compounds (CVOCs) are persistent pollutants and harmful to the atmosphere, environment, and human health. The catalytic elimination of CVOCs has become a hotspot of interest due to their self-toxicity, the secondary generation of chlorinated by-products, and the Cl poisoning of [...] Read more.
Chlorinated volatile organic compounds (CVOCs) are persistent pollutants and harmful to the atmosphere, environment, and human health. The catalytic elimination of CVOCs has become a hotspot of interest due to their self-toxicity, the secondary generation of chlorinated by-products, and the Cl poisoning of catalysts. The development of high-performance, highly selective, and anti-poisoning catalysts is a critical issue. Bimetallic catalysts exhibit an improved dechlorination performance, poisoning resistance, and product selectivity through the modulation of geometrical and electronic structures. The present review article gives a brief overview of the recent advancements in the preparation of bimetallic catalysts and their catalytic CVOC elimination activities. In addition, representative case studies are provided to investigate the physicochemical properties, CVOC conversion, COx and inorganic chlorine species selectivities, and by-product control so that the structure–performance relationships of bimetallic catalysts can be established. Furthermore, this review article provides a fundamental understanding of designing bimetallic catalysts with specific active components and the desired physicochemical properties for target reactions. In the end, related perspectives for future work are proposed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Possible oxidation mechanism of dichloromethane over the Ni–V/TiO<sub>2</sub> catalyst [<a href="#B47-catalysts-14-00531" class="html-bibr">47</a>]. Reprinted with permission from Ref. [<a href="#B46-catalysts-14-00531" class="html-bibr">46</a>]. Copyright 2019, Elsevier.</p>
Full article ">Figure 2
<p>Adsorption model and adsorption energy of 4-chlorophenol and phenol on the (111) crystal plane of pure Pd, Pd<sub>3</sub>Au<sub>1</sub>, Pd<sub>2</sub>Au<sub>2</sub> and pure Au, a Pd atom (gray), a Au atom (yellow), a C atom (brown), a H atom (white), an O atom (red), and a Cl atom (green) [<a href="#B58-catalysts-14-00531" class="html-bibr">58</a>]. Reprinted with permission from Ref. [<a href="#B57-catalysts-14-00531" class="html-bibr">57</a>]. Copyright 2022, Elsevier.</p>
Full article ">Figure 3
<p>(<b>a</b>) The reaction pathway for electrocatalytic hydrodechlorination of 2,4-DCP on the AgPd bimetallic nanoparticle (NP) catalysts, and (<b>b</b>) a schematic illustration of the role of Ag for electrocatalytic hydrodechlorination in AgPd NP catalysts [<a href="#B58-catalysts-14-00531" class="html-bibr">58</a>]. Reprinted with permission from Ref. [<a href="#B58-catalysts-14-00531" class="html-bibr">58</a>]. Copyright 2019, American Chemical Society.</p>
Full article ">Figure 4
<p>The possible dechlorination pathways of PCBs based on (<b>a</b>) the bond dissociation energy (kcal/mol), (<b>b</b>) the C–Cl bond length, and (<b>c</b>) the lowest unoccupied molecular orbital (LUMO) of PCB anions [<a href="#B59-catalysts-14-00531" class="html-bibr">59</a>]. Reprinted with permission from Ref. [<a href="#B59-catalysts-14-00531" class="html-bibr">59</a>]. Copyright 2023, Elsevier.</p>
Full article ">Figure 5
<p>DFT and schematic diagram of Pd@CoFe<sub>V</sub>-LDH. (<b>a</b>) Optimized adsorption configurations and adsorption energies of 2,4-dichlorophenol; (<b>b</b>) schematic profiles for energy reaction barriers of water activation pathway (Volmer Step) on Pd@CoFe<sub>V</sub>-LDH, CoFe<sub>V</sub>-LDH, and Pd NPs; (<b>c</b>) free-energy diagram for HER; and (<b>d</b>) schematic diagram of facilitation for EHDC reaction of 2,4-dichlorophenol over Pd@CoFe<sub>V</sub>-LDH/NF catalyst [<a href="#B61-catalysts-14-00531" class="html-bibr">61</a>]. Reprinted with permission from Ref. [<a href="#B61-catalysts-14-00531" class="html-bibr">61</a>]. Copyright 2024, Elsevier.</p>
Full article ">Figure 6
<p>(<b>a</b>) The optimal H* adsorption configuration for Pd, Pd/Cu, Pd/CuO, and Pd/Cu<sub>2</sub>O; (<b>b</b>) the corresponding free energy diagram for the HER over Pd, Pd/Cu, Pd/CuO, and Pd/Cu<sub>2</sub>O; and (<b>c</b>) a schematic diagram of synergistic interaction over Pd/CuO<span class="html-italic"><sub>x</sub></span> [<a href="#B62-catalysts-14-00531" class="html-bibr">62</a>]. Reprinted with permission from Ref. [<a href="#B62-catalysts-14-00531" class="html-bibr">62</a>]. Copyright 2024, Elsevier.</p>
Full article ">
15 pages, 4943 KiB  
Article
Dual-Function Photocatalysis in the Visible Spectrum: Ag-G-TiO2 for Simultaneous Dye Wastewater Degradation and Hydrogen Production
by Tarek Ahasan, Pei Xu and Huiyao Wang
Catalysts 2024, 14(8), 530; https://doi.org/10.3390/catal14080530 - 15 Aug 2024
Viewed by 468
Abstract
Photocatalytic processes offer promising solutions for environmental remediation and clean energy production, yet their efficiency under the visible light spectrum remains a significant challenge. Here, we report a novel silver–graphene (Ag-G) modified TiO2 (Ag-G-TiO2) nanocomposite photocatalyst that demonstrates remarkably enhanced [...] Read more.
Photocatalytic processes offer promising solutions for environmental remediation and clean energy production, yet their efficiency under the visible light spectrum remains a significant challenge. Here, we report a novel silver–graphene (Ag-G) modified TiO2 (Ag-G-TiO2) nanocomposite photocatalyst that demonstrates remarkably enhanced photocatalytic activity for both dye wastewater degradation and hydrogen production under visible and UV light irradiation. Through comprehensive characterization and performance analysis, we reveal that the Ag-G modification narrows the TiO2 bandgap from 3.12 eV to 1.79 eV, enabling efficient visible light absorption. The nanocomposite achieves a peak hydrogen production rate of 191 μmolesg−1h−1 in deionized (DI) water dye solution under visible light, significantly outperforming unmodified TiO2. Intriguingly, we observe an inverse relationship between dye degradation efficiency and hydrogen production rates in dye solutions with tap water versus DI water, highlighting the critical role of water composition in photocatalytic processes. This work not only advances the understanding of fundamental photocatalytic mechanisms but also presents a promising photocatalyst for solar-driven environmental remediation and clean energy production. The Ag-G-TiO2 nanocomposite’s enhanced performance across both visible and UV spectra, coupled with its dual functionality in dye degradation and hydrogen evolution, represents a significant step towards addressing critical challenges in water treatment and sustainable energy generation. Our findings highlight the complex interplay between light absorption and reaction conditions, offering new insights for optimizing photocatalytic systems. This research paves the way for developing more efficient and versatile photocatalysts, potentially contributing to the global transition towards sustainable technologies and circular economy in waste management and energy production. Full article
(This article belongs to the Special Issue Advances in Photocatalytic Wastewater Purification, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>TEM images of the catalysts and their particle size distributions: (<b>A</b>) pure TiO<sub>2</sub>, (<b>B</b>) Ag-G-TiO<sub>2</sub>, (<b>C</b>) 200K magnification of Ag-G-TiO<sub>2</sub>, (<b>D</b>) pure TiO<sub>2</sub>, (<b>E</b>) Ag-G-TiO<sub>2</sub>.</p>
Full article ">Figure 1 Cont.
<p>TEM images of the catalysts and their particle size distributions: (<b>A</b>) pure TiO<sub>2</sub>, (<b>B</b>) Ag-G-TiO<sub>2</sub>, (<b>C</b>) 200K magnification of Ag-G-TiO<sub>2</sub>, (<b>D</b>) pure TiO<sub>2</sub>, (<b>E</b>) Ag-G-TiO<sub>2</sub>.</p>
Full article ">Figure 2
<p>(<b>A</b>) EDS analysis of Ag-G-TiO<sub>2</sub>, (<b>B</b>) weight percentage bar chart of the elements.</p>
Full article ">Figure 3
<p>XPS spectra of pure TiO<sub>2</sub> and Ag-G-TiO<sub>2</sub>.</p>
Full article ">Figure 4
<p>XPS spectra of (<b>A</b>) Ti 2p, (<b>B</b>) O 1s, (<b>C</b>) C 1s, (<b>D</b>) Ag 3d, (<b>E</b>) Elemental weight percentage.</p>
Full article ">Figure 4 Cont.
<p>XPS spectra of (<b>A</b>) Ti 2p, (<b>B</b>) O 1s, (<b>C</b>) C 1s, (<b>D</b>) Ag 3d, (<b>E</b>) Elemental weight percentage.</p>
Full article ">Figure 5
<p>Bandgap analysis of the catalysts: (<b>A</b>) pure TiO<sub>2</sub>, (<b>B</b>) Ag-G-TiO<sub>2</sub>, (<b>C</b>) Absorbance comparison between pure TiO<sub>2</sub> and Ag-G-TiO<sub>2</sub>.</p>
Full article ">Figure 6
<p>XRD patterns of the pure TiO<sub>2</sub> and Ag-G-TiO<sub>2</sub>.</p>
Full article ">Figure 7
<p>Adsorption isotherm of Ag-G-TiO<sub>2</sub> catalyst.</p>
Full article ">Figure 8
<p>Photocatalytic dye wastewater degradation and catalyst efficiency over time. (<b>A</b>) Visible light, (<b>B</b>) UV light, (<b>C</b>) visible light, (<b>D</b>) UV light.</p>
Full article ">Figure 9
<p>Photocatalytic hydrogen production. (<b>A</b>) Visible light, (<b>B</b>) UV light.</p>
Full article ">
31 pages, 9696 KiB  
Article
Production of Green Fuel Using a New Synthetic Magnetite Mesoporous Nano-Silica Composite Catalyst for Oxidative Desulfurization: Experiments and Process Modeling
by Aysar T. Jarullah, Ahmed K. Hussein, Ban A. Al-Tabbakh, Shymaa A. Hameed, Iqbal M. Mujtaba, Liqaa I. Saeed and Jasim I. Humadi
Catalysts 2024, 14(8), 529; https://doi.org/10.3390/catal14080529 - 15 Aug 2024
Viewed by 461
Abstract
Producing an eco-friendly fuel with the least amount of sulfur compounds has been an ongoing issue for petroleum refineries. In this study, bentonite (which is a cheap material and is locally available in abundance) is employed to prepare a nano-silica catalyst with a [...] Read more.
Producing an eco-friendly fuel with the least amount of sulfur compounds has been an ongoing issue for petroleum refineries. In this study, bentonite (which is a cheap material and is locally available in abundance) is employed to prepare a nano-silica catalyst with a high surface area to be used for the oxidative desulfurization of kerosene. Two composite catalysts of Fe/silica were supported on CAT-1 (0% HY-zeolite and 100% nano-silica) and CAT-2 (20% HY-zeolite and 80% nano-silica). The activity of the catalysts was evaluated in a batch ODS (oxidative desulfurization) process at temperatures of 30, 60, 90, and 120 °C, a pressure of 1 atm, and a reaction time of 30, 60, 90, and 120 min using 120 L/h of air as the oxidant. The results revealed that the highest total sulfur removal efficiency was 50% and 87.88% for 100% nano-silica (CAT-1) and 80% nano-silica (CAT-2), respectively. The experimental data were then used to construct and validate an accurate mathematical model of the process. The operational parameters for eliminating more than 99% of sulfur and producing eco-friendly fuel were then achieved by using the model. The testing methods for these characterizing materials included X-ray diffraction (XRD), thermal gravimetric examination (TGA), X-ray fluorescence (XRF), and surface area (BET). The outcomes indicated that the addition of HY-zeolite increased the activity of the catalyst (CAT-2 > CAT-1). Full article
Show Figures

Figure 1

Figure 1
<p>XRD for amorphous nano-silica.</p>
Full article ">Figure 2
<p>Analysis of nano-silica using FTIR.</p>
Full article ">Figure 2 Cont.
<p>Analysis of nano-silica using FTIR.</p>
Full article ">Figure 3
<p>FTIR spectrum for CAT-2.</p>
Full article ">Figure 4
<p>TGA analysis of (<b>A</b>) nano-silica and (<b>B</b>) composite catalyst.</p>
Full article ">Figure 5
<p>The size distributions of nano-silica and the produced catalysts.</p>
Full article ">Figure 6
<p>SEM image of nano-silica.</p>
Full article ">Figure 7
<p>EDX image of nano-silica.</p>
Full article ">Figure 8
<p>FESEM of the catalyst.</p>
Full article ">Figure 9
<p>EDX mapping of the catalyst.</p>
Full article ">Figure 10
<p>Effect of reaction temperature on sulfur conversion at several reaction times. Effect of reaction temperature on sulfur conversion for different catalyst types at various times: (<b>A</b>) 30 min, (<b>B</b>) 60 min, (<b>C</b>) 90 min, (<b>D</b>) 120 min.</p>
Full article ">Figure 11
<p>Effect of reaction time for different catalyst types on the sulfur conversion at various reaction temperatures. Effect of reaction time on sulfur conversion at different temperatures: (<b>A</b>) 303 K, (<b>B</b>) 333 K, (<b>C</b>) 363 K, (<b>D</b>) 393 K.</p>
Full article ">Figure 12
<p>CAT-1 oxidation kinetics: ln(k) vs. 1/T.</p>
Full article ">Figure 13
<p>Kinetics of oxidation with CAT-2, displaying lnk against 1/T.</p>
Full article ">Figure 14
<p>Comparison of CAT-1 data from experiments and simulations.</p>
Full article ">Figure 15
<p>Comparison of CAT-2 data from experiments and simulations.</p>
Full article ">Figure 16
<p>Steps of nano-silica preparation.</p>
Full article ">Figure 17
<p>Catalyst preparation using the IWI method.</p>
Full article ">Figure 18
<p>Schematic layout of a batch ODS reactor.</p>
Full article ">
12 pages, 5341 KiB  
Article
Defective TiO2/MIL-88B(Fe) Photocatalyst for Tetracycline Degradation: Characterization and Augmented Photocatalytic Efficiency
by Dongsheng Xiang, Zhihao Wang, Jingwen Xu, Hongdan Shen, Xiaodong Zhang and Ning Liu
Catalysts 2024, 14(8), 528; https://doi.org/10.3390/catal14080528 - 15 Aug 2024
Viewed by 442
Abstract
Photocatalysts, such as TiO2, are widely used in photoreduction. However, drawbacks like their wide band gap and short carrier lifetime lead to lower efficiencies with their use. Introducing defects and forming heterostructures of TiO2 could extend the carrier’s light-harvesting range [...] Read more.
Photocatalysts, such as TiO2, are widely used in photoreduction. However, drawbacks like their wide band gap and short carrier lifetime lead to lower efficiencies with their use. Introducing defects and forming heterostructures of TiO2 could extend the carrier’s light-harvesting range from UV to visible light and enhance its lifetime. Herein, an electron-beam irradiation-defected TiO2 was induced in MIL-88B(Fe). The structure of the material was characterized using XRD, FT-IR, TEM, HRTEM, and XPS techniques. Remarkably, TiO2 under 300 kGy electron-beam irradiation performed the best with a series of 0, 100, 300, and 500 kGy irradiation ratios. PL and UV–vis DRS were utilized to measure the material’s optical properties. The introduction of MIL-88B(Fe) expanded the light response range, reduced the optical band gap, and lengthened the carrier lifetime of the defective TiO2 composite photocatalysts, resulting in superior TC photoreduction capabilities of 88B5%300, which degraded 97% of tetracycline (10 mg/L) in water after 120 min. Full article
(This article belongs to the Special Issue Green Chemistry and Catalysis)
Show Figures

Figure 1

Figure 1
<p>88B@5%300 characterizations: (<b>a</b>,<b>b</b>) XRD, (<b>c</b>,<b>d</b>) FT-IR.</p>
Full article ">Figure 2
<p>XPS of 88B@5%300: (<b>a</b>) survey, (<b>b</b>) C 1 s, (<b>c</b>) O 1 s, (<b>d</b>) Fe 2 p, (<b>e</b>) Ti 2 p.</p>
Full article ">Figure 3
<p>(<b>a</b>) TEM and (<b>b</b>) HRTEM of 88B@5%300 (<b>b1</b>–<b>b4</b>) Different effects of electron beam irradiation on the lattice).</p>
Full article ">Figure 4
<p>Photocatalytic experiments of TiO<sub>2</sub> composite catalysts: (<b>a</b>) irradiation dosages, (<b>b</b>) composite amounts of defective TiO<sub>2</sub>, synergistic effects, (<b>c</b>) catalyst amount and (<b>d</b>) PDS concentration.</p>
Full article ">Figure 5
<p>(<b>a</b>) Effect of initial pH values and (<b>b</b>) Recycle experiment of 88B@5%300 (reaction condition: [88B@5%300]<sub>0</sub> = 5 mg, [PDS]<sub>0</sub> = 500 ppm]).</p>
Full article ">Figure 6
<p>(<b>a</b>) Results of quenching experiments, and (<b>b</b>) UV–vis DRS spectrum, (<b>c</b>) Tauc-Plot, (<b>d</b>) PL spectra of 88B@5%300 and MIL-88B (reaction condition: [88B@5%300]<sub>0</sub> = 5 mg, [PDS]<sub>0</sub> = 500 ppm]).</p>
Full article ">Figure 7
<p>Schema of photocatalytic mechanism.</p>
Full article ">Figure 8
<p>Schema of preparation procedure of 88B@X%YYY.</p>
Full article ">
13 pages, 4063 KiB  
Article
A Highly Efficient Tribocatalysis of La/ZnO Powders for Degradation of Rhodamine B
by Dobrina K. Ivanova, Bozhidar I. Stefanov and Nina V. Kaneva
Catalysts 2024, 14(8), 527; https://doi.org/10.3390/catal14080527 - 15 Aug 2024
Viewed by 372
Abstract
Tribocatalysis is a promising environmental remediation technique that utilizes the triboelectric effect, produced when dissimilar materials interact through friction, to generate charges promoting catalytic reactions. In this work, the tribocatalytic degradation of an organic dye—Rhodamine B (RhB)—has been experimentally realized using pure and [...] Read more.
Tribocatalysis is a promising environmental remediation technique that utilizes the triboelectric effect, produced when dissimilar materials interact through friction, to generate charges promoting catalytic reactions. In this work, the tribocatalytic degradation of an organic dye—Rhodamine B (RhB)—has been experimentally realized using pure and 2 mol.% La-modified/ZnO powders, synthesized via a simple hydrothermal method. The effects of annealing on the tribocatalytic activity of the La/ZnO catalysts are also studied at 100 and 500 °C. The La/ZnO-modified catalysts showed an enhanced RhB degradation efficiency with 92% removal within 24 h, compared to only 58% for the pure ZnO. The effects of annealing were found to be detrimental, with RhB removal efficiencies dropping from 92 to 69% in the 100–500 °C range. The catalysts’ cycling stability was found to be excellent within three cycles. Ultimately, it is demonstrated that by utilizing La/ZnO powders, contaminated wastewater can be efficiently treated through employing tribocatalysis. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>SEM images of powders annealed at different temperatures: (<b>a</b>) ZnO, 100 °C; (<b>b</b>) La/ZnO, 100 °C; (<b>c</b>) La/ZnO, 200 °C; (<b>d</b>) La/ZnO, 300 °C; (<b>e</b>) La/ZnO, 400 °C; and (<b>f</b>) La/ZnO, 500 °C.</p>
Full article ">Figure 2
<p>BET adsorption isotherm data obtained for: (<b>a</b>) pure ZnO, (<b>b</b>) La/ZnO, 100 °C.</p>
Full article ">Figure 3
<p>XRD pattern of ZnO and La/ZnO powders annealed at 100, 300 and 500 °C.</p>
Full article ">Figure 4
<p>TEM micrographs from selected areas of samples La/ZnO powders annealed at 100 °C at: (<b>a</b>) low and (<b>b</b>) high magnification.</p>
Full article ">Figure 5
<p>(<b>a</b>) Stirring degradation of Rhodamine B solution using ZnO and La/ZnO powders annealed at 100 °C by magnetic stirring conditions, 500 rpm; (<b>b</b>) kinetic fitting.</p>
Full article ">Figure 6
<p>Effect of scavengers on Rhodamine B degradation in tribocatalysis process using (<b>a</b>) pure and (<b>b</b>) lanthanum-modified ZnO powder.</p>
Full article ">Figure 7
<p>(<b>a</b>) Tribocatalytic decomposition of Rhodamine B by La/ZnO powders annealed at different annealed temperatures; (<b>b</b>) kinetic fitting; (<b>c</b>) degrees of RhB dye degradation versus preparation temperature of La-modified ZnO samples.</p>
Full article ">Figure 8
<p>ATR-FTIR spectra obtained for the 0th, 12th and 24th hour of Rhodamine B tribodegradation for: (<b>a</b>) pure ZnO; (<b>b</b>) La/ZnO 100 °C; (<b>c</b>) La/ZnO 200 °C; (<b>d</b>) La/ZnO 300 °C; (<b>e</b>) La/ZnO 400 °C; (<b>f</b>) La/ZnO 500 °C.</p>
Full article ">Figure 9
<p>Tribocatalytic degradation rate of Rhodamine B for three consecutive cycles using La/ZnO powders, annealed at different temperatures.</p>
Full article ">Figure 10
<p>Plausible tribocatalytic degradation pathway of Rhodamine B. Mechanism adapted with permission from Ref. [<a href="#B45-catalysts-14-00527" class="html-bibr">45</a>].</p>
Full article ">
24 pages, 5842 KiB  
Article
Porous Nanostructured Catalysts Based on Silicates and Their Surface Functionality: Effects of Silica Source and Metal Added in Glycerol Valorization
by José Vitor C. Carmo, Joabson Nogueira, Gabriela M. Bertoldo, Francisco E. Clemente, Alcineia C. Oliveira, Adriana F. Campos, Gian C. S. Duarte, Samuel Tehuacanero-Cuapa, José Jiménez-Jiménez and Enrique Rodríguez-Castellón
Catalysts 2024, 14(8), 526; https://doi.org/10.3390/catal14080526 - 15 Aug 2024
Viewed by 452
Abstract
A series of nanospherical-shaped silicates containing heteroatoms (Al, Zr or Ti) were successfully synthesized using tetraethylorthosilicate (TEOS) or silica colloids as a silicon source. These metallosilicate nanospheres were used as silicon nutrients to obtain silicalite zeolites with micro-mesoporosity and improved textural properties. The [...] Read more.
A series of nanospherical-shaped silicates containing heteroatoms (Al, Zr or Ti) were successfully synthesized using tetraethylorthosilicate (TEOS) or silica colloids as a silicon source. These metallosilicate nanospheres were used as silicon nutrients to obtain silicalite zeolites with micro-mesoporosity and improved textural properties. The results demonstrated that TEOS acted as a suitable silicon source to produce amorphous silicates and a spherical-type zeolite architecture with Zr and Ti heteroatoms included in their framework, with preferable particle size and crystallinity. The surface functionality of the mesostructured nanospheres and zeolite silicates provide active centers for the esterification of glycerol with acetic acid (EG) reaction. The dispersion of Cu entities on the surface of the zeolites achieved high glycerol conversions selectively producing triacetin in comparison with Fe counterparts. Full article
(This article belongs to the Special Issue Novel Nanocatalysts for Sustainable and Green Chemistry)
Show Figures

Figure 1

Figure 1
<p>SEM micrographs, EDS mappings and EDS spectra of the as-synthesized silica-based nanospheres obtained by using the following silica sources: (<b>a</b>) TEOS and (<b>b</b>) colloidal silica. The included figures are the magnified regions of the images.</p>
Full article ">Figure 1 Cont.
<p>SEM micrographs, EDS mappings and EDS spectra of the as-synthesized silica-based nanospheres obtained by using the following silica sources: (<b>a</b>) TEOS and (<b>b</b>) colloidal silica. The included figures are the magnified regions of the images.</p>
Full article ">Figure 2
<p>SEM micrographs of calcined samples: (<b>a</b>) silica-based mesostructured spheres, and (<b>b</b>) as-synthesized silica-seeded mesostructured zeolites.</p>
Full article ">Figure 2 Cont.
<p>SEM micrographs of calcined samples: (<b>a</b>) silica-based mesostructured spheres, and (<b>b</b>) as-synthesized silica-seeded mesostructured zeolites.</p>
Full article ">Figure 3
<p>(<b>a</b>) Nitrogen physisorption isotherms and (<b>b</b>) pore-size distributions of the as-synthesized silica-based mesostructured spheres.</p>
Full article ">Figure 4
<p>XRD patterns of the solids in study: (<b>a</b>) as-synthesized silica-based mesostructured spheres, (<b>b</b>) calcined silica-based mesostructured spheres and (<b>c</b>) calcined silica-seeded mesostructured zeolites.</p>
Full article ">Figure 5
<p>FTIR spectra of the (<b>a</b>) as-synthesized silica-based mesostructured spheres, (<b>b</b>) calcined silica-based mesostructured spheres and (<b>c</b>) calcined silica-seeded mesostructured zeolites. The included figure is the FTIR spectrum of NZ-SAT sample.</p>
Full article ">Figure 6
<p>Representative XPS spectra of selected solids: (<b>a</b>) XPS survey spectrum, (<b>b</b>) Si 2<span class="html-italic">p</span>, (<b>c</b>) O 1<span class="html-italic">s</span>, (<b>d</b>) Na 1<span class="html-italic">s</span>, (<b>e</b>) C 1<span class="html-italic">s</span> and (<b>f</b>) Al 2<span class="html-italic">p</span> core-level spectra for NZ-SAT. The XPS spectra of the STT-C and SZT-C samples corresponding to the (<b>g</b>) Ti 2<span class="html-italic">p</span> and (<b>h</b>) Zr 3<span class="html-italic">d</span> core levels.</p>
Full article ">Figure 6 Cont.
<p>Representative XPS spectra of selected solids: (<b>a</b>) XPS survey spectrum, (<b>b</b>) Si 2<span class="html-italic">p</span>, (<b>c</b>) O 1<span class="html-italic">s</span>, (<b>d</b>) Na 1<span class="html-italic">s</span>, (<b>e</b>) C 1<span class="html-italic">s</span> and (<b>f</b>) Al 2<span class="html-italic">p</span> core-level spectra for NZ-SAT. The XPS spectra of the STT-C and SZT-C samples corresponding to the (<b>g</b>) Ti 2<span class="html-italic">p</span> and (<b>h</b>) Zr 3<span class="html-italic">d</span> core levels.</p>
Full article ">Figure 7
<p>The catalytic performance of the (<b>a</b>) solids in the EG reaction, and (<b>b</b>) effects of Fe and Cu on the catalytic performance of the most active solids.</p>
Full article ">Figure 8
<p>Schematic representation of the synthetic route used to produce the obtained solids: (<b>a</b>) as-synthesized silica-based mesostructured spheres prepared from TEOS and colloidal silica sources, (<b>b</b>) silica-seeded mesostructured zeolites synthesis.</p>
Full article ">
20 pages, 4448 KiB  
Article
Biogenic Synthesis Based on Cuprous Oxide Nanoparticles Using Eucalyptus globulus Extracts and Its Effectiveness for Removal of Recalcitrant Compounds
by Pablo Salgado, Katherine Márquez and Gladys Vidal
Catalysts 2024, 14(8), 525; https://doi.org/10.3390/catal14080525 - 14 Aug 2024
Viewed by 603
Abstract
Recalcitrant compounds resulting from anthropogenic activity are a significant environmental challenge, necessitating the development of advanced oxidation processes (AOPs) for effective remediation. This study explores the synthesis of cuprous oxide nanoparticles on cellulose-based paper (Cu2O@CBP) using Eucalyptus globulus leaf extracts, leveraging [...] Read more.
Recalcitrant compounds resulting from anthropogenic activity are a significant environmental challenge, necessitating the development of advanced oxidation processes (AOPs) for effective remediation. This study explores the synthesis of cuprous oxide nanoparticles on cellulose-based paper (Cu2O@CBP) using Eucalyptus globulus leaf extracts, leveraging green synthesis techniques. The scanning electron microscopy (SEM) analysis found the average particle size 64.90 ± 16.76 nm, X-ray diffraction (XRD) and Raman spectroscopy confirm the Cu2O structure in nanoparticles; Fourier-transform infrared spectroscopy (FTIR) suggests the reducing role of phenolic compounds; and ultraviolet–visible diffuse reflectance spectroscopy (UV-Vis DRS) allowed us to determine the band gap (2.73 eV), the energies of the valence band (2.19 eV), and the conduction band (−0.54 eV) of Cu2O@CBP. The synthesized Cu2O catalysts demonstrated efficient degradation of methylene blue (MB) used as a model as recalcitrant compounds under LED-driven visible light photocatalysis and heterogeneous Fenton-like reactions with hydrogen peroxide (H2O2) using the degradation percentage and the first-order apparent degradation rate constant (kapp). The degradation efficiency of MB was pH-dependent, with neutral pH favoring photocatalysis (kapp = 0.00718 min−1) due to enhanced hydroxyl (·OH) and superoxide radical (O2·) production, while acidic pH conditions improved Fenton-like reaction efficiency (kapp = 0.00812 min−1) via ·OH. The reusability of the photocatalysts was also evaluated, showing a decline in performance for Fenton-like reactions at acidic pH about 22.76% after five cycles, while for photocatalysis at neutral pH decline about 11.44% after five cycles. This research provides valuable insights into the catalytic mechanisms and supports the potential of eco-friendly Cu2O nanoparticles for sustainable wastewater treatment applications. Full article
(This article belongs to the Section Photocatalysis)
Show Figures

Figure 1

Figure 1
<p>Photographic images of (<b>a</b>) pristine cellulose-based paper and (<b>b</b>) Cu<sub>2</sub>O@CBP for 4 × 4 cm pieces.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) pristine paper, (<b>b</b>) Cu<sub>2</sub>O@CBP, and (<b>c</b>) particle size distribution of Cu<sub>2</sub>O@CBP (The red line represents the best curve fitting using the Gaussian distribution function).</p>
Full article ">Figure 3
<p>(<b>a</b>) Raman spectra for Cu<sub>2</sub>O@CBP (red numbers for Cu<sub>2</sub>O nanoparticles, green numbers for phenolic compounds, and black numbers for cellulose), (<b>b</b>) XRD patterns of pristine cellulose-based paper and Cu<sub>2</sub>O@CBP.</p>
Full article ">Figure 4
<p>FTIR analyses for pristine cellulose-based paper and Cu<sub>2</sub>O@CBP.</p>
Full article ">Figure 5
<p>(<b>a</b>) Band gap energy (E<sub>g</sub>) determination from the Tauc plot (green line) for Cu<sub>2</sub>O@CBP. The linear part of the plot is extrapolated to the x-axis using a dotted line for E<sub>g</sub> determination. (inset graph: DRS spectra for Cu<sub>2</sub>O@CBP). (<b>b</b>) Potential band diagram for Cu<sub>2</sub>O@CBP.</p>
Full article ">Figure 6
<p>Analysis of (<b>a</b>) total phenolic compounds, (<b>b</b>) total reducing sugars, (<b>c</b>) total proteins, (<b>d</b>) FRAP, and (<b>e</b>) CUPRAC in <span class="html-italic">E. globulus</span> extract before and after the synthesis of Cu<sub>2</sub>O@CBP. Differences between groups were compared using ANOVA with Tukey post hoc analysis. **** <span class="html-italic">p</span> &lt; 0.0001, ns: no significant differences.</p>
Full article ">Figure 7
<p>FTIR spectra of <span class="html-italic">E. globulus</span> extract before (black spectra) and after synthesis (green spectra) of Cu<sub>2</sub>O@CBP.</p>
Full article ">Figure 8
<p>MB removal efficiency by (<b>a</b>) photocatalysis under LED visible light and (<b>b</b>) Fenton-like reaction catalyzed by Cu<sub>2</sub>O@CBP at pH 3.0 and 7.0 (inset: legend represents samples at different pH). (<b>c</b>) Photocatalytic and (<b>d</b>) Fenton-like calculated degradation rate constant (<span class="html-italic">k<sub>app</sub></span>) for Cu<sub>2</sub>O@CBP at pH 3.0 and 7.0 (inset: legend represents samples at different pH, <span class="html-italic">k<sub>app</sub></span>, and r<sup>2</sup> of a pseudo-first order model).</p>
Full article ">Figure 9
<p>Cycling performance of Cu<sub>2</sub>O@CBP for MB degradation by (<b>a</b>) photocatalysis under LED visible light at pH = 7.0 and (<b>b</b>) Fenton-like at pH = 3.0 under dark. The data obtained are presented as mean ± standard deviation.</p>
Full article ">Figure 10
<p>MB removal by Cu<sub>2</sub>O@CBP under influence of scavenging agents (<b>a</b>) Na<sub>2</sub>C<sub>2</sub>O<sub>4</sub>, CrO<sub>3</sub>, BQ, and IP on photocatalytic degradation under LED visible light, and (<b>b</b>) BQ and IP on Fenton-like under darkness. The data obtained are presented as mean ± standard deviation.</p>
Full article ">Figure 11
<p>Scheme of Fenton-like reaction at acidic pH under dark and photocatalysis at neutral pH under LED visible light to generate reactive species and MB degradation process.</p>
Full article ">Figure 12
<p>(<b>a</b>) Internal view and (<b>b</b>) view from above of the reactor used for photocatalysis and Fenton-like reaction degradation of MB by Cu<sub>2</sub>O@CBP.</p>
Full article ">
4 pages, 171 KiB  
Editorial
Special Issue Catalysis for Bitumen/Heavy Oil Upgrading and Petroleum Refining
by Irek I. Mukhamatdinov and Nikita N. Sviridenko
Catalysts 2024, 14(8), 524; https://doi.org/10.3390/catal14080524 - 13 Aug 2024
Viewed by 500
Abstract
Currently, fossil fuels continue to play a crucial role in daily life [...] Full article
(This article belongs to the Special Issue Catalysis for Bitumen/Heavy Oil Upgrading and Petroleum Refining)
Previous Issue
Back to TopTop