[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (181)

Search Parameters:
Keywords = catalytic ozonation

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 3121 KiB  
Article
Effect of the Fe2O3/SBA-15 Surface on Inducing Ozone Decomposition and Mass Transfer in Water
by Lei Yuan, Lele Fang, Jizhou Zhang, Pengwei Yan and Zhonglin Chen
Water 2024, 16(18), 2590; https://doi.org/10.3390/w16182590 - 12 Sep 2024
Viewed by 246
Abstract
Catalytic ozonation with metal oxides is of interest for advanced water treatment technology. The amount of active oxygen-containing radicals produced is a primary objective of this process. Fe2O3 is a widely used catalyst because of its high performance. In this [...] Read more.
Catalytic ozonation with metal oxides is of interest for advanced water treatment technology. The amount of active oxygen-containing radicals produced is a primary objective of this process. Fe2O3 is a widely used catalyst because of its high performance. In this study, Fe2O3/SBA-15 was synthesized and characterized. The results revealed that Fe2O3/SBA-15 was a nano-/mesoporous material with high-order hexagonal array structures and exhibited greater catalytic performance than Fe2O3 in ozonation processes. To investigate the role of the Fe2O3/SBA-15 surface in O3 decomposition, the kinetic constant was measured, and the interfacial reactions were discussed. Compared with Fe2O3, Fe2O3/SBA-15 significantly increased the formation of hydroxyl radicals (•OH) and the efficient utilization of O3 in the catalytic O3 decomposition process. The SBA-15 support decreased the O3 self-decomposition rate during catalytic ozonation with Fe2O3/SBA-15, which resulted in increased formation of •OH via the reaction between O3 and Fe2O3. From a practical point of view, Fe2O3/SBA-15 is an efficient green ozonation catalyst for water treatment. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the experimental apparatus.</p>
Full article ">Figure 2
<p>(<b>A</b>) Wide-angle XRD patterns, (<b>B</b>) low-angle XRD patterns, (<b>C</b>) physisorption isotherms, (<b>D</b>) pore size distributions, (<b>E</b>) FT-IR spectra, and (<b>F</b>) SEM image of Fe<sub>2</sub>O<sub>3</sub>/SBA-15.</p>
Full article ">Figure 3
<p>Removal of nitrobenzene in different processes (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of nitrobenzene in solution: 50 μg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; concentration of SBA-15: 45 mg/L; initial pH of solution: 6.0; reaction time: 10 min; reaction; temperature: 25 °C; and concentration of TBA: 2 mg/L).</p>
Full article ">Figure 4
<p>O<sub>3</sub> utilization efficiency in different processes (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of nitrobenzene in solution: 50 μg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; initial pH of the solution: 6.0; reaction time: 10 min; reaction; and temperature: 25 °C).</p>
Full article ">Figure 5
<p>Influence of water temperature on removal (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of nitrobenzene in solution: 50 μg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; initial pH of the solution: 6.0; and reaction time: 10 min).</p>
Full article ">Figure 6
<p>Influence of catalyst content on removal (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of nitrobenzene in solution: 50 μg/L; initial pH of the solution: 6.0; reaction time: 10 min; and reaction temperature: 25 °C).</p>
Full article ">Figure 7
<p>(<b>A</b>) Ozone decay and (<b>B</b>) pseudo-first-order plot in water (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; initial pH of the solution: 6.0; and reaction temperature: 25 °C).</p>
Full article ">Figure 8
<p>Influence of pH on (<b>A</b>) ozone decay and (<b>B</b>) pseudo-first-order plot (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; and reaction temperature: 25 °C).</p>
Full article ">Figure 9
<p>•OH captured via ESR (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; initial pH of the solution: 6.0; reaction temperature: 25 °C; and concentration of DMPO: 100 mmol/L).</p>
Full article ">Figure 10
<p>Ions released into water (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; initial pH of the solution: 6.0; reaction temperature: 25 °C; and reaction time: 30 min).</p>
Full article ">Figure 11
<p>•OH captured via ESR (test conditions: concentration of O<sub>3</sub> in solution: 0.6 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>/SBA-15: 50 mg/L; concentration of Fe<sub>2</sub>O<sub>3</sub>: 5 mg/L; initial pH of the solution: 6.0; reaction temperature: 25 °C; and concentration of DMPO: 100 mmol/L).</p>
Full article ">
15 pages, 3424 KiB  
Article
Carbon-Based Materials in Combined Adsorption/Ozonation for Indigo Dye Decolorization in Constrain Contact Time
by Naghmeh Fallah, Ermelinda Bloise, Elisa I. García-López and Giuseppe Mele
Molecules 2024, 29(17), 4144; https://doi.org/10.3390/molecules29174144 - 31 Aug 2024
Viewed by 391
Abstract
This study presents a comprehensive evaluation of catalytic ozonation as an effective strategy for indigo dye bleaching, particularly examining the performance of four carbon-based catalysts, activated carbon (AC), multi-walled carbon nanotubes (MWCNT), graphitic carbon nitride (g-C3N4), and thermally etched [...] Read more.
This study presents a comprehensive evaluation of catalytic ozonation as an effective strategy for indigo dye bleaching, particularly examining the performance of four carbon-based catalysts, activated carbon (AC), multi-walled carbon nanotubes (MWCNT), graphitic carbon nitride (g-C3N4), and thermally etched nanosheets (C3N4-TE). The study investigates the efficiency of catalytic ozonation in degrading Potassium indigotrisulfonate (ITS) dye within the constraints of short contact times, aiming to simulate real-world industrial wastewater treatment conditions. The results reveal that all catalysts demonstrated remarkable decolorization efficiency, with over 99% of indigo dye removed within just 120 s of mixing time. Besides, the study delves into the mechanisms underlying catalytic ozonation reactions, elucidating the intricate interactions between the catalysts, ozone, and indigo dye molecules with the processes being influenced by factors such as PZC, pKa, and pH. Furthermore, experiments were conducted to analyze the adsorption characteristics of indigo dye on the surfaces of the materials and its impact on the catalytic ozonation process. MWCNT demonstrated the highest adsorption efficiency, effectively removing 43.4% of the indigo dye color over 60 s. Although the efficiency achieved with C3N4-TE was 21.4%, which is approximately half of that achieved with MWCNT and less than half of that with AC, it is noteworthy given the significantly lower surface area of C3N4-TE. Full article
(This article belongs to the Special Issue Recent Research Progress of Novel Ion Adsorbents)
Show Figures

Figure 1

Figure 1
<p>FTIR spectra of AC, MWCNT, g-C<sub>3</sub>N<sub>4</sub>, and C<sub>3</sub>N<sub>4</sub>-TE.</p>
Full article ">Figure 2
<p>SEM image of AC (<b>a</b>); MWCNT (<b>b</b>); g-C<sub>3</sub>N<sub>4</sub> (<b>c</b>); and C<sub>3</sub>N<sub>4</sub>-TE (<b>d</b>).</p>
Full article ">Figure 3
<p>XRD pattern of AC (<b>a</b>); MWCNT (<b>b</b>); g-C<sub>3</sub>N<sub>4</sub> (<b>c</b>); and C<sub>3</sub>N<sub>4</sub>-TE (<b>d</b>).</p>
Full article ">Figure 4
<p>Decolorization efficiency of ITS over time using different doses of AC (<b>a</b>), MWCNT (<b>b</b>), g-C<sub>3</sub>N<sub>4</sub> (<b>c</b>), and C<sub>3</sub>N<sub>4</sub>-TE (<b>d</b>).</p>
Full article ">Figure 5
<p>Comparing the decolorization efficiency of ITS over time using AC, MWCNT, g-C<sub>3</sub>N<sub>4</sub>, and C<sub>3</sub>N<sub>4</sub>-TE (1.5 g/L).</p>
Full article ">Figure 6
<p>Comparing the catalytic ozonation efficiency for ITS removal over time using 0.25 g/L AC, MWCNT, g-C<sub>3</sub>N<sub>4</sub>, and C<sub>3</sub>N<sub>4</sub>-TE.</p>
Full article ">Figure 7
<p>Catalytic ozonation mechanisms proposed for AC (<b>a</b>); MWCNT (<b>b</b>); g-C<sub>3</sub>N<sub>4</sub> and C<sub>3</sub>N<sub>4</sub>-TE (<b>c</b>).</p>
Full article ">Figure 8
<p>Comparing the catalytic ozonation efficiency for ITS decolorization over time using (<b>a</b>) AC, (<b>b</b>) MWCNT, (<b>c</b>) g-C<sub>3</sub>N<sub>4</sub>, and (<b>d</b>) C<sub>3</sub>N<sub>4</sub>-TE.</p>
Full article ">
12 pages, 4555 KiB  
Article
Boosting Benzene’s Ozone Catalytic Oxidation at Mild Temperatures over Highly Dispersed Ag-Doped Mn3O4
by Hao Guo, Liwei Cen, Kui Deng, Wenlong Mo, Hojo Hajime, Di Hu, Pan Zhang, Wenfeng Shangguan, Haibao Huang and Hisahiro Einaga
Catalysts 2024, 14(9), 554; https://doi.org/10.3390/catal14090554 - 23 Aug 2024
Viewed by 378
Abstract
Transition metal oxides show high activity while still facing the challenges of low mineralization and poor durability in the ozone catalytic oxidation (OCO) of volatile organic compounds (VOCs). Improving the oxygen mobility and low-temperature reducibility of transition metal oxides was found to be [...] Read more.
Transition metal oxides show high activity while still facing the challenges of low mineralization and poor durability in the ozone catalytic oxidation (OCO) of volatile organic compounds (VOCs). Improving the oxygen mobility and low-temperature reducibility of transition metal oxides was found to be an effective way to address the above challenges. Here, highly dispersed Ag was added to Mn3O4 via the co-precipitation oxalate route, and the obtained Ag/Mn3O4 exhibited higher mineralization and stability in benzene catalytic ozonation at room temperature. Compared to Mn3O4, the concentration of CO2 formed from benzene oxidation over Ag/Mn3O4 was significantly increased, from 585.4 ppm to 810.9 ppm, while CO generation was greatly suppressed to only one tenth of its original value (194 ppm vs. 19 ppm). In addition, Ag/Mn3O4 exhibited higher catalytic stability than Mn3O4. The introduction of Ag obviously improved the oxygen mobility and low-temperature reducibility of Mn3O4. Moreover, the highly dispersed Ag also promoted the activity of surface oxygen species and the chemisorption of benzene on Mn3O4. The above physicochemical properties contributed to the excellent catalytic performance and durability of Ag/Mn3O4. This research could shed light on the improvement in VOC mineralization via ozone catalytic oxidation. Full article
(This article belongs to the Special Issue Catalytic Energy Conversion and Catalytic Environmental Purification)
Show Figures

Figure 1

Figure 1
<p>XRD spectra (<b>a</b>) and N<sub>2</sub>-adsorption/desorption isotherms (<b>b</b>), inset pore size distribution curves) of the obtained catalysts.</p>
Full article ">Figure 2
<p>HRTEM image (<b>a</b>), HAADF-STEM image (<b>b</b>), and elemental mapping ((<b>c</b>): Mn, (<b>d</b>): Ag) of Ag/Mn<sub>3</sub>O<sub>4</sub>.</p>
Full article ">Figure 3
<p>XPS curves of Mn<sub>3</sub>O<sub>4</sub> and Ag/Mn<sub>3</sub>O<sub>4</sub> ((<b>a</b>): Mn 2p 3/2; (<b>b</b>): Ag 2d 5/2; and (<b>c</b>): O 1 s).</p>
Full article ">Figure 4
<p>O<sub>2</sub>-TPD (<b>a</b>) and H<sub>2</sub>-TPR (<b>b</b>) profiles of as-prepared catalysts.</p>
Full article ">Figure 5
<p>C<sub>6</sub>H<sub>6</sub>-TPSR profiles of Mn<sub>3</sub>O<sub>4</sub> (<b>a</b>) and Ag/Mn<sub>3</sub>O<sub>4</sub> (<b>b</b>).</p>
Full article ">Figure 6
<p>Benzene conversion and concentrations of CO<sub>2</sub>, CO, and O<sub>3</sub> as a function of time at 25 °C and 70 °C, respectively ((<b>a,b</b>): Mn<sub>3</sub>O<sub>4</sub>; (<b>c</b>,<b>d</b>): Ag/Mn<sub>3</sub>O<sub>4</sub>)<b>.</b></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 458
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, 2641 KiB  
Article
From Waste to Resource: Evaluating Biomass Residues as Ozone-Catalyst Precursors for the Removal of Recalcitrant Water Pollutants
by Cátia A. L. Graça and Olívia Salomé Gonçalves Pinto Soares
Environments 2024, 11(8), 172; https://doi.org/10.3390/environments11080172 - 12 Aug 2024
Viewed by 709
Abstract
Five different biomass wastes—orange peel, coffee grounds, cork, almond shell, and peanut shell—were transformed into biochars (BCs) or activated carbons (ACs) to serve as adsorbents and/or ozone catalysts for the removal of recalcitrant water treatment products. Oxalic acid (OXL) was used as a [...] Read more.
Five different biomass wastes—orange peel, coffee grounds, cork, almond shell, and peanut shell—were transformed into biochars (BCs) or activated carbons (ACs) to serve as adsorbents and/or ozone catalysts for the removal of recalcitrant water treatment products. Oxalic acid (OXL) was used as a model pollutant due to its known refractory character towards ozone. The obtained materials were characterized by different techniques, namely thermogravimetric analysis, specific surface area measurement by nitrogen adsorption, and elemental analysis. In adsorption experiments, BCs generally outperformed ACs, except for cork-derived materials. Orange peel BC revealed the highest adsorption capacity (Qe = 40 mg g−1), while almond shell BC showed the best cost–benefit ratio at €0.0096 per mg of OXL adsorbed. In terms of catalytic ozonation, only ACs made from cork and coffee grounds presented significant catalytic activity, achieving pollutant removal rates of 72 and 64%, respectively. Among these materials, ACs made from coffee grounds reveal the best cost/benefit ratio with €0.02 per mg of OXL degraded. Despite the cost analysis showing that these materials are not the cheapest options, other aspects rather than the price alone must be considered in the decision-making process for implementation. This study highlights the promising role of biomass wastes as precursors for efficient and eco-friendly water treatment processes, whether as adsorbents following ozone water treatment or as catalysts in the ozonation reaction itself. Full article
(This article belongs to the Special Issue Advanced Research on Micropollutants in Water)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Adsorption capacities Qe (mg g<sup>−1</sup>).</p>
Full article ">Figure 2
<p>OXL degradation profile promoted by BCs combined with O<sub>3</sub>. [OXL]<sub>0</sub> = 20 mg·L<sup>−</sup><sup>1</sup>, BCs = 50 mg·L<sup>−</sup><sup>1</sup>, [O<sub>3</sub>]<sub>gas</sub> = 50 g·Nm<sup>3</sup>.</p>
Full article ">Figure 3
<p>OXL degradation profile promoted by ACs combined with O<sub>3</sub>. [OXL]<sub>0</sub> = 20 mg·L<sup>−1</sup>, ACs = 50 mg·L<sup>−1</sup>, [O<sub>3</sub>]<sub>gas</sub> = 50 g·Nm<sup>3</sup>.</p>
Full article ">Figure A1
<p>TGA and DTG curves for orange peel.</p>
Full article ">Figure A2
<p>TGA and DTG curves for coffee ground.</p>
Full article ">Figure A3
<p>TGA and DTG curves for cork waste.</p>
Full article ">Figure A4
<p>TGA and DTG curves for almond shell.</p>
Full article ">Figure A5
<p>TGA and DTG curves for peanut shell.</p>
Full article ">
31 pages, 14363 KiB  
Article
Hybrid Dielectric Barrier Discharge Reactor: Characterization for Ozone Production
by Dariusz Korzec, Florian Freund, Christian Bäuml, Patrik Penzkofer and Stefan Nettesheim
Plasma 2024, 7(3), 585-615; https://doi.org/10.3390/plasma7030031 - 27 Jul 2024
Viewed by 569
Abstract
The generation of ozone by dielectric barrier discharge (DBD) is widely used for water and wastewater treatment, the control of catalytic reactions, and surface treatment. Recently, a need for compact, effective, and economical ozone and reactive oxygen–nitrogen species (RONS) generators for medical, biological, [...] Read more.
The generation of ozone by dielectric barrier discharge (DBD) is widely used for water and wastewater treatment, the control of catalytic reactions, and surface treatment. Recently, a need for compact, effective, and economical ozone and reactive oxygen–nitrogen species (RONS) generators for medical, biological, and agricultural applications has been observed. In this study, a novel hybrid DBD (HDBD) reactor fulfilling such requirements is presented. Its structured high-voltage (HV) electrode allows for the ignition of both the surface and volume microdischarges contributing to plasma generation. A Peltier module cooling of the dielectric barrier, made of alumina, allows for the efficient control of plasma chemistry. The typical electrical power consumption of this device is below 30 W. The operation frequency of the DBD driver oscillating in the auto-resonance mode is from 20 to 40 kHz. The specific energy input (SEI) of the reactor was controlled by the DBD driver input voltage in the range from 10.5 to 18.0 V, the Peltier current from 0 to 4.5 A, the duty cycle of the pulse-width modulated (PWM) power varied from 0 to 100%, and the gas flow from 0.5 to 10 SLM. The operation with oxygen, synthetic air, and compressed dry air (CDA) was characterized. The ultraviolet light (UV) absorption technique was implemented for the measurement of the ozone concentration. The higher harmonics of the discharge current observed in the frequency range of 5 to 50 MHz were used for monitoring the discharge net power. Full article
(This article belongs to the Special Issue Processes in Atmospheric Pressure Plasmas)
Show Figures

Figure 1

Figure 1
<p>Setup for HDBD reactor characterization.</p>
Full article ">Figure 2
<p>The schematic cross-sectional view of the HDBD reactor used for ozone generation.</p>
Full article ">Figure 3
<p>The hybrid SDBD-VDBD discharge operation principle. (<b>a</b>) Visualization by use of a glass plate coated with ITO, placed at a tilt on the HV electrode surface. (<b>b</b>) The volume microdischarge in the gap between the post surface and the dielectric barrier. (<b>c</b>) The hybrid DBD with surface and volume microdischarges at the post touching the dielectric barrier surface.</p>
Full article ">Figure 4
<p>(<b>a</b>) Two PWM cycles of PWM, and (<b>b</b>) two cycles of kHz excitation of the high voltage measured between the HDBD electrodes as a function of time for the driver input voltage of 12 V, PWM frequency of 100 Hz, PWM duty cycle of 40%, CDA flow of 1 SLM, and Peltier module current of 2 A.</p>
Full article ">Figure 5
<p>The output RMS high voltage and apparent power for load capacity and resistance of (1) 2 pF and 1 M<math display="inline"><semantics> <mi mathvariant="sans-serif">Ω</mi> </semantics></math> (triangle), (2) 40 pF and 150 k<math display="inline"><semantics> <mi mathvariant="sans-serif">Ω</mi> </semantics></math> (square), and (3) 80 pF and 300 k<math display="inline"><semantics> <mi mathvariant="sans-serif">Ω</mi> </semantics></math> (circle), respectively, as a function of the input DC voltage of the DBD driver.</p>
Full article ">Figure 6
<p>The current spectrum measured for the DBD operating in air at different power densities.</p>
Full article ">Figure 7
<p>Influence of the oxygen gas flow on the ozone concentration expressed in ppm (<b>a</b>) and MDIR signal compared with ozone production rate (<b>b</b>) for the duty cycle of 100%, the Peltier current of 0 A, and with three DBD driver input voltages, as depicted at the curves.</p>
Full article ">Figure 8
<p>The ozone concentration and ozone production rate in pure oxygen, shown as a function of drive voltage for the duty cycle of 100%, the Peltier current of 0 A, and with four oxygen flows in SLM, as depicted at the curves.</p>
Full article ">Figure 9
<p>The ozone concentration expressed in volume percentage (<b>a</b>) and MDIR signal (<b>b</b>) shown as a function of the duty cycle of PWM for HDBD reactor operated with pure oxygen, switched off Peltier cooling, 0.6 SLM oxygen flow, and three driver input voltages, as depicted at the curves.</p>
Full article ">Figure 10
<p>The ozone concentration in pure oxygen expressed in volume percentage (<b>a</b>) and MDIR signal (<b>b</b>), shown as a function of the Peltier module current for the HDBD reactor operated without pulse-width modulation, with 0.6 SLM oxygen flow, and with three driver input voltages as labeled at the curves. The fitting functions used for the sensitivity calculation in Equation (<a href="#FD11-plasma-07-00031" class="html-disp-formula">11</a>) are included.</p>
Full article ">Figure 11
<p>The ozone concentration and ozone production rate are shown as a function of synthetic air flow for the duty cycle of 80%, the Peltier current of 0 A, and three DBD driver input voltages, as depicted in the diagram.</p>
Full article ">Figure 12
<p>Influence of the duty cycle on ozone production rate at the synthetic air flow of (<b>a</b>) 0.6 SLM, and (<b>b</b>) 10 SLM and on MDIR signal voltage at synthetic air flow of (<b>c</b>) 0.6 SLM, and (<b>d</b>) 10 SLM with DBD driver input voltage as a parameter, and Peltier current of 0 A.</p>
Full article ">Figure 13
<p>The ozone concentration for the DBD driver input voltage of (<b>a</b>) 10.5 V, and (<b>b</b>) 15 V, and the MDIR signal for the DBD driver input voltage of (<b>c</b>) 10.5 V, and (<b>d</b>) 15 V, shown as a function of the Peltier module current for the HDBD reactor operated with the duty cycle of 80%, for synthetic air flow varying from 0.6 to 10 SLM, as depicted at the curves.</p>
Full article ">Figure 14
<p>The ozone concentration and ozone production rate, shown as a function of CDA flow for the duty cycle of 80%, the Peltier current of 0 A, and with three DBD driver input voltages, as depicted in the curves.</p>
Full article ">Figure 15
<p>The ozone production rate as a function of duty cycle with DBD driver input voltage as a parameter; Peltier current of 0 A, compared for four CDA flows: (<b>a</b>) 0.6 SLM, (<b>b</b>) 1.0 SLM, (<b>c</b>) 5.0 SLM, (<b>d</b>) 10.0 SLM.</p>
Full article ">Figure 16
<p>The ozone concentration is shown as a function of the Peltier module current for the HDBD reactor operated with the duty cycle of 80%, the DBD driver input voltage of (<b>a</b>) 10.5 V, and (<b>b</b>) 15 V, for CDA flow varying from 0.6 to 10 SLM, as depicted at the curves.</p>
Full article ">Figure 17
<p>Influence of the Peltier module current on the ozone concentration at the CDA flow (<b>a</b>) 0.6 SLM and (<b>b</b>) 2.0 SLM, and on the MDIR signal at the CDA flow of (<b>c</b>) 0.6 SLM and (<b>d</b>) 2.0 SLM as a function of duty cycle for the DBD driver input voltage of 10.5 V.</p>
Full article ">Figure 18
<p>The limiting lines, separating the regions of the effective and ineffective Peltier cooling, in the CDA flow vs. duty cycle coordinate system for three driver voltages.</p>
Full article ">
14 pages, 3355 KiB  
Article
Kinetics and H2O Influence on NOx Trapping and Selective Catalytic Reduction over Ce/Pd Doping Catalyst
by Li Yang and Tianshan Xue
Molecules 2024, 29(15), 3457; https://doi.org/10.3390/molecules29153457 - 24 Jul 2024
Viewed by 458
Abstract
In this paper, the removal effects and activation energy of Ce and Pd doping on pollutants (CO, C3H6, and NO) were comparatively analyzed by using characterization methods and constructed kinetic equations. Furthermore, the problems of the water influence mechanism [...] Read more.
In this paper, the removal effects and activation energy of Ce and Pd doping on pollutants (CO, C3H6, and NO) were comparatively analyzed by using characterization methods and constructed kinetic equations. Furthermore, the problems of the water influence mechanism on the NSR process were also discussed. The results show the following: (1) Pd doping effectively improves the removal of CO (80%) and C3H6 (71%) in the low-temperature section of the catalyst (150–250 °C) compared to Ce doping, while Ce doping exhibits excellent low-temperature conversion of NO. (2) The reaction activation energy of the LaKMnPdO3 catalyst was 9784 kJ/mol, which was significantly lower than that of the LaKMnCeO3 catalyst. (3) The presence of H2O has an important enhancement effect in the storage performance of the LaKMnPdO3 catalyst for NOx but decreases the catalytic reduction of NO. It provides a solution for the effective treatment of the increasing problems of particulate matter and ozone pollution. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction patterns for catalysts; (<b>b</b>) FT-IR of LaKMnO<sub>3</sub>, LaKMnCeO<sub>3</sub>, and LaKMnPdO<sub>3</sub> catalysts.</p>
Full article ">Figure 2
<p>SEM images of catalysts: (<b>a</b>) LaKMnO<sub>3</sub>; (<b>b</b>) LaKMnCeO<sub>3</sub>; (<b>c</b>) LaKMnPdO<sub>3</sub>.</p>
Full article ">Figure 3
<p>C<sub>3</sub>H<sub>6</sub>, CO, and NO conversion for catalysts. (<b>a</b>) CO conversion; (<b>b</b>) C<sub>3</sub>H<sub>6</sub> conversion; (<b>c</b>) NO conversion.</p>
Full article ">Figure 4
<p>XPS of catalysts ((<b>a</b>) LaKMnO<sub>3</sub>, (<b>b</b>) LaKMnCeO<sub>3</sub>, and (<b>c</b>) LaKMnPdO<sub>3</sub>).</p>
Full article ">Figure 5
<p>Results of De-NO<sub>x</sub> process performed for catalysts ((<b>a</b>) LaKMnCeO<sub>3</sub>; (<b>b</b>) LaKMnPdO<sub>3</sub>).</p>
Full article ">Figure 6
<p>Fitting curves of lgr<sub>NO</sub>, lgC<sub>NO</sub> and lgr<sub>CO</sub>, lgC<sub>CO</sub>.</p>
Full article ">Figure 7
<p>Fitting curves of lgr<sub>NO</sub>, lgC<sub>NO</sub> and lgr<sub>CO</sub>, lgC<sub>CO</sub>.</p>
Full article ">Figure 8
<p>Catalytic performance of LaKMnPdO<sub>3</sub> for reduction of C<sub>3</sub>H<sub>6</sub>, NO, and N<sub>2</sub> yielsd under different water volumes.</p>
Full article ">Figure 9
<p>The FT-IR of LaKMnPdO<sub>3</sub> with different active component loadings: (<b>a</b>) before the experiment; (<b>b</b>) after the experiment; and (<b>c</b>) in the presence of water.</p>
Full article ">Figure 10
<p>Alternate experimental results of rich and lean combustion of LaKMnPdO<sub>3</sub> catalyst under water-free and water conditions.</p>
Full article ">
25 pages, 4111 KiB  
Review
Global Trends in the Research and Development of Petrochemical Waste Gas from 1981 to 2022
by Mengting Wu, Wei Liu, Zhifei Ma, Tian Qin, Zhiqin Chen, Yalan Zhang, Ning Cao, Xianchuan Xie, Sunlin Chi, Jinying Xu and Yi Qi
Sustainability 2024, 16(14), 5972; https://doi.org/10.3390/su16145972 - 12 Jul 2024
Viewed by 790
Abstract
As a highly energy-intensive and carbon-emitting industry with significant emissions of volatile organic compounds (VOCs), the petroleum and chemical industry is a major contributor to the global greenhouse effect and ozone layer destruction. Improper treatment of petrochemical waste gas (PWG) seriously harms human [...] Read more.
As a highly energy-intensive and carbon-emitting industry with significant emissions of volatile organic compounds (VOCs), the petroleum and chemical industry is a major contributor to the global greenhouse effect and ozone layer destruction. Improper treatment of petrochemical waste gas (PWG) seriously harms human health and the natural environment. This study uses CiteSpace and VOSviewer to conduct a scientometric analysis of 1384 scholarly works on PWG and carbon sequestration published between 1981 and 2022, revealing the basic characteristics, knowledge base, research topic evolution, and research hotspots of the field. The results show the following: (1) In the early stages of the petrochemical industry, it was processed tail gas, plant leakage waste gas, and combustion flue gas that were investigated in PWG research. (2) Later, green environmental protection technology was widely studied in the field of PWG treatment, such as biotechnology, catalytic oxidation technology, membrane separation technology, etc., in order to achieve efficient, low energy consumption and low emissions of waste gas treatment, and the number of publications related to this topic has increased rapidly. In addition, researchers studied the internet of things and technology integration, such as the introduction of artificial intelligence, big data analysis, and other technologies, to improve the accuracy and efficiency of exhaust gas monitoring, control, and management. (3) The department has focused on how to reduce emissions by optimizing petrochemical process lines or improving energy efficiency. Emission reduction and low-carbon transition in the petrochemical industry will become the main trend in the future. Switching from renewable carbon to feedstock carbon derived from captured carbon dioxide, biomass, or recycled chemicals has become an attractive strategy to help curb emissions from the chemical industry. The results of our analysis can provide funding agencies and research groups with information to better understand the global trends and directions that have emerged in this field from 1981 to 2022 and serve as a reference for future research. Full article
Show Figures

Figure 1

Figure 1
<p>Annual publication volume, global trend of publications, and type of publications (1981–2022).</p>
Full article ">Figure 2
<p>Publications in more than 15 institutions (<b>a</b>) and journals (<b>b</b>).</p>
Full article ">Figure 3
<p>Country cooperation network in PWG (1981–2022).</p>
Full article ">Figure 4
<p>Institutional cooperation network in PWG (1981–2022).</p>
Full article ">Figure 5
<p>Authors’ cooperation network in PWG (1981–2022).</p>
Full article ">Figure 6
<p>Co-citation reference network (1981–2022) and corresponding cluster analysis. Note: Co-citation reference network with cluster visualization and hotspot outburst.</p>
Full article ">Figure 7
<p>(<b>a</b>) VOS-based keyword co-occurrence. (<b>b</b>) Timeline visualization of co-occurring keyword networks (1981–2022).</p>
Full article ">Figure 8
<p>Top 25 keywords with the strongest citation bursts. Sorted by burst intensity (1993–2022). The beginning of a blue line represents when an article is published. The beginning of a red mark represents the beginning of a period of burst, and the end of the red mark is the end of the burst period (1993–2022).</p>
Full article ">
47 pages, 26240 KiB  
Review
The Structures and Compositions Design of the Hollow Micro–Nano-Structured Metal Oxides for Environmental Catalysis
by Jingxin Xu, Yufang Bian, Wenxin Tian, Chao Pan, Cai-e Wu, Leilei Xu, Mei Wu and Mindong Chen
Nanomaterials 2024, 14(14), 1190; https://doi.org/10.3390/nano14141190 - 12 Jul 2024
Viewed by 694
Abstract
In recent decades, with the rapid development of the inorganic synthesis and the increasing discharge of pollutants in the process of industrialization, hollow-structured metal oxides (HSMOs) have taken on a striking role in the field of environmental catalysis. This is all due to [...] Read more.
In recent decades, with the rapid development of the inorganic synthesis and the increasing discharge of pollutants in the process of industrialization, hollow-structured metal oxides (HSMOs) have taken on a striking role in the field of environmental catalysis. This is all due to their unique structural characteristics compared to solid nanoparticles, such as high loading capacity, superior pore permeability, high specific surface area, abundant inner void space, and low density. Although the HSMOs with different morphologies have been reviewed and prospected in the aspect of synthesis strategies and potential applications, there has been no systematic review focusing on the structures and compositions design of HSMOs in the field of environmental catalysis so far. Therefore, this review will mainly focus on the component dependence and controllable structure of HSMOs in the catalytic elimination of different environmental pollutants, including the automobile and stationary source emissions, volatile organic compounds, greenhouse gases, ozone-depleting substances, and other potential pollutants. Moreover, we comprehensively reviewed the applications of the catalysts with hollow structure that are mainly composed of metal oxides such as CeO2, MnOx, CuOx, Co3O4, ZrO2, ZnO, Al3O4, In2O3, NiO, and Fe3O4 in automobile and stationary source emission control, volatile organic compounds emission control, and the conversion of greenhouse gases and ozone-depleting substances. The structure–activity relationship is also briefly discussed. Finally, further challenges and development trends of HSMO catalysts in environmental catalysis are also prospected. Full article
(This article belongs to the Collection Metallic and Metal Oxide Nanohybrids and Their Applications)
Show Figures

Figure 1

Figure 1
<p>Schematic of the reaction mechanism for CO oxidation. Figure reproduced form ref. [<a href="#B71-nanomaterials-14-01190" class="html-bibr">71</a>].</p>
Full article ">Figure 2
<p>The SEM (x1 and x2) and TEM (x3) images of the Cu<sub>2</sub>O cubes, composite CeO<sub>2</sub>-Cu<sub>2</sub>O (1), NiO@Cu<sub>2</sub>O (2), and CeO<sub>2</sub>–NiO–Cu<sub>2</sub>O (3). (x = a, b, c and d for Cu<sub>2</sub>O and composites 1−3, respectively) The scale bar is 800 nm in parts x1 and x3, and it is 500 nm in part x2. Figure reproduced from ref. [<a href="#B89-nanomaterials-14-01190" class="html-bibr">89</a>]. (<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>) The SEM images of Cu<sub>2</sub>O cubes; (<b>a<sub>3</sub></b>) The TEM images of Cu<sub>2</sub>O cubes; (<b>a<sub>4</sub></b>) The SAED pattern of Cu<sub>2</sub>O cubes; (<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>) The SEM images of composite CeO<sub>2</sub>-Cu<sub>2</sub>O; (<b>b<sub>3</sub></b>) The TEM images of composite CeO<sub>2</sub>-Cu<sub>2</sub>O; (<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>) The SEM images of composite NiO@Cu<sub>2</sub>O; (<b>c<sub>3</sub></b>) The TEM images of composite NiO@Cu<sub>2</sub>O; (<b>d<sub>1</sub></b>,<b>d<sub>2</sub></b>) The SEM images of composite CeO<sub>2</sub>–NiO–Cu<sub>2</sub>O; (<b>d<sub>3</sub></b>) The TEM images of composite CeO<sub>2</sub>–NiO–Cu<sub>2</sub>O.</p>
Full article ">Figure 3
<p>(<b>a</b>) The formation process of the hollow Pd–CeO<sub>2</sub> nano-composite sphere reproduced from ref. [<a href="#B91-nanomaterials-14-01190" class="html-bibr">91</a>]; (<b>b</b>) Schematic illustration of the synthesis process for sandwich-like MnO<sub>2</sub>–Pd–CeO<sub>2</sub> hollow spheres. Figure reproduced from ref. [<a href="#B93-nanomaterials-14-01190" class="html-bibr">93</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) The reaction mechanism of CO oxidation over the In<sub>2</sub>O<sub>3</sub>@Pd–Co<sub>3</sub>O<sub>4</sub> catalyst reproduced from ref. [<a href="#B108-nanomaterials-14-01190" class="html-bibr">108</a>]; (<b>b</b>,<b>c</b>) SEM images, (<b>d</b>–<b>f</b>) TEM images, and (<b>g</b>) HRTEM images of the MnO<sub>2</sub>–Co<sub>3</sub>O<sub>4</sub> hollow spheres reproduced from ref. [<a href="#B109-nanomaterials-14-01190" class="html-bibr">109</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>f</b>) SEM images of the MnO<sub>x</sub> with hollow morphology; (<b>g</b>) Reaction pathways of NO oxidation over H-MnO<sub>2</sub> and MnO<sub>2</sub>-R catalysts reproduced from ref. [<a href="#B113-nanomaterials-14-01190" class="html-bibr">113</a>].</p>
Full article ">Figure 6
<p>TEM images of the CeO<sub>2</sub>–MnO<sub>x</sub> hollow spheres with various shell numbers obtained at different heating rates: (<b>a</b>) before calcination; (<b>b</b>) single-shell, 2 °C min<sup>−1</sup>; (<b>c</b>) double-shell, 5 °C min<sup>−1</sup>; (<b>d</b>) triple-shell, 10 °C min<sup>−1</sup>. Insets show the corresponding individual hollow sphere; (<b>e</b>) Proposed collision processes of reactive gases against hollow spheres with different shells. Figure reproduced from ref. [<a href="#B122-nanomaterials-14-01190" class="html-bibr">122</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>,<b>c</b>,<b>e</b>) SEM images of the CMC-Cp-a, -b , and -c, respectively; (<b>b</b>,<b>d</b>,<b>f</b>) TEM images of the CMC-Cp-a, -b, and -c, respectively (Ce/Mn/Cu molar ratio of 84/16/16 (CMC-Cp-a), 84/8/8 (CMC-Cp-b) or 84/32/16 (CMC-Cp-c)) reproduced from ref. [<a href="#B136-nanomaterials-14-01190" class="html-bibr">136</a>]; (<b>g</b>,<b>h</b>) TEM images of the CeO<sub>2</sub>@MnO<sub>2</sub> reproduced from ref. [<a href="#B139-nanomaterials-14-01190" class="html-bibr">139</a>].</p>
Full article ">Figure 8
<p>Schematic diagram of the complete reaction cycle for the catalytic oxidation of toluene on MnO<sub>2</sub>-1.2. Figure reproduced from ref. [<a href="#B151-nanomaterials-14-01190" class="html-bibr">151</a>].</p>
Full article ">Figure 9
<p>SEM and HRTEM images of: (<b>a</b>–<b>c</b>) solid-urchin and (<b>d</b>–<b>f</b>) hollow-urchin MnO<sub>2</sub>. Plausible reaction pathways for toluene decomposition in the PPC (post-plasma catalysis) process are also given: (<b>g</b>) NTP (non-thermal plasma)-induced gas-phase reactions in the DBD (dielectric barrier discharge) reactor and (<b>h</b>) catalytic reactions on the surface of MnO<sub>2</sub> in the catalytic reactor. Figure reproduced from ref. [<a href="#B159-nanomaterials-14-01190" class="html-bibr">159</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) The schematic illustration of the reaction mechanism for toluene oxidation over Pd/metal oxide catalysts reproduced from ref. [<a href="#B6-nanomaterials-14-01190" class="html-bibr">6</a>]; (<b>b</b>) Proposed mechanism for enhanced catalytic oxidation toward toluene over 2.0 wt% Pt/Co<sub>2.73</sub>Zr<sub>0.27</sub>O<sub>4</sub> reproduced from ref. [<a href="#B20-nanomaterials-14-01190" class="html-bibr">20</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>) Illustration of two pathways to synthesize the Cu–Mn oxide reproduced from ref. [<a href="#B155-nanomaterials-14-01190" class="html-bibr">155</a>]; (<b>b</b>) Schematic of the oxidation of toluene on Mn<sub>x</sub>Co<sub>3−x</sub>O<sub>4</sub> reproduced from ref. [<a href="#B158-nanomaterials-14-01190" class="html-bibr">158</a>]; (<b>c</b>) The formation schematic of the porous hollow HC-CoInO<sub>x</sub> nanocube; (<b>d</b>) The proposed reaction mechanism over the CoInO<sub>x</sub> catalyst reproduced from ref. [<a href="#B150-nanomaterials-14-01190" class="html-bibr">150</a>].</p>
Full article ">Figure 12
<p>Brief description of the hard templating method to synthesize the MnO<sub>2</sub> hollow sphere. Figure reproduced from ref. [<a href="#B23-nanomaterials-14-01190" class="html-bibr">23</a>].</p>
Full article ">Figure 13
<p>FESEM images of the MnO<sub>2</sub>-HCS (<b>a</b>,<b>b</b>). TEM ((<b>c</b>) and inset of (<b>d</b>)) and HRTEM (<b>d</b>) images of the Pt/MnO<sub>2</sub>-HCS reproduced from ref. [<a href="#B2-nanomaterials-14-01190" class="html-bibr">2</a>]; (<b>e</b>–<b>h</b>) Structural characterization of the Pt/γ-Al<sub>2</sub>O<sub>3</sub>; SEM (<b>e</b>), high magnification SEM (inset in (<b>e</b>)), TEM (<b>f</b>,<b>g</b>), and HRTEM (<b>h</b>) images of the PHAO sample reproduced from ref. [<a href="#B169-nanomaterials-14-01190" class="html-bibr">169</a>]; (<b>i</b>–<b>l</b>) SEM images and the corresponding high-magnification SEM images (insets) of the samples: Ni80 (<b>i</b>), Ni400 (<b>j</b>), and Ni600 (<b>k</b>); TEM image of the Ni400P sample (<b>l</b>) reproduced from ref. [<a href="#B170-nanomaterials-14-01190" class="html-bibr">170</a>].</p>
Full article ">Figure 14
<p>The SEM and TEM images of the prepared Fe<sub>2</sub>O<sub>3</sub> (<b>a</b>) and Fe<sub>2</sub>O<sub>3</sub>@SnO<sub>2</sub> (<b>b</b>–<b>d</b>) reproduced from ref. [<a href="#B172-nanomaterials-14-01190" class="html-bibr">172</a>].</p>
Full article ">Figure 15
<p>(<b>a</b>,<b>b</b>) FESEM and (<b>c</b>) TEM images of the synthesized CuAl<sub>2</sub>O<sub>4</sub> hollow sphere reproduced from ref. [<a href="#B179-nanomaterials-14-01190" class="html-bibr">179</a>]; (<b>d</b>–<b>f</b>) SEM images at different magnifications of the CuO-ZnO catalyst reproduced from ref. [<a href="#B183-nanomaterials-14-01190" class="html-bibr">183</a>]. The red borders in the figure represent irregular spherical shape of the sample.</p>
Full article ">Figure 16
<p>TEM images of the Fe<sub>3</sub>O<sub>4</sub>–Au 5 mL (<b>a</b>) and Fe<sub>3</sub>O<sub>4</sub>–Au 20 mL with the HRTEM image (inset) (<b>b</b>); Fe<sub>3</sub>O<sub>4</sub>–Au 40 mL with the SAED pattern (inset) (<b>c</b>) and Fe<sub>3</sub>O<sub>4</sub>–Au 60 mL (<b>d</b>); TEM images of single Fe<sub>3</sub>O<sub>4</sub>–Au 5 mL (<b>e</b>) and Fe<sub>3</sub>O<sub>4</sub>–Au 60 mL (<b>f</b>) microspheres and corresponding EDS elemental mapping images (Au, Fe, and O). Figure reproduced from ref. [<a href="#B201-nanomaterials-14-01190" class="html-bibr">201</a>] TEM images of (<b>g</b>) hollow Fe<sub>3</sub>O<sub>4</sub> microspheres, (<b>h</b>) Fe<sub>3</sub>O<sub>4</sub>/P(GMA-EGDMA) microspheres, and (<b>i</b>,<b>j</b>) Fe<sub>3</sub>O<sub>4</sub>/P (GMA-EGDMA)SO<sub>3</sub>H/Au-PPy microspheres. Figure reproduced from ref. [<a href="#B202-nanomaterials-14-01190" class="html-bibr">202</a>]. TEM images of (<b>k</b>) SiO<sub>2</sub> nanospheres, (<b>l</b>) Au/Fe<sub>3</sub>O<sub>4</sub>/SiO<sub>2</sub>, (<b>m</b>) Au/Fe<sub>3</sub>O<sub>4</sub>/SiO<sub>2</sub>@TiO<sub>2</sub>, (<b>n</b>) Au/Fe<sub>3</sub>O<sub>4</sub>@hTiO<sub>2</sub>. Figure reproduced from ref. [<a href="#B203-nanomaterials-14-01190" class="html-bibr">203</a>].</p>
Full article ">Figure 17
<p>(<b>a</b>) Schematic diagram of the catalytic mechanism of metal oxide HNSs@C for the hydrogenation of 4-NP reaction. Figure reproduced from ref. [<a href="#B207-nanomaterials-14-01190" class="html-bibr">207</a>]. SEM images of (<b>b</b>) Fe<sub>2</sub>O<sub>3</sub>, (<b>c</b>) FeCa5, (<b>d</b>) FeCa10, and (<b>e</b>) FeCa20. Figure reproduced from ref. [<a href="#B209-nanomaterials-14-01190" class="html-bibr">209</a>]. SEM images of (<b>f</b>) FeMn10, (<b>g</b>) FeMn20, (<b>h</b>) FeMn40, and (<b>i</b>) FeMn80 (represent the molar ratios of Mn/(Fe + Mn) are ~10, 20, 40, and 80 mol%). Figure reproduced from ref. [<a href="#B22-nanomaterials-14-01190" class="html-bibr">22</a>]. (<b>j</b>) Proposed reaction routes of <span class="html-italic">o</span>-DCB catalytic oxidation over ZnCe5 (doped with 5 mol% Ce). Figure reproduced from ref. [<a href="#B210-nanomaterials-14-01190" class="html-bibr">210</a>].</p>
Full article ">Figure 18
<p>SEM images of the (<b>a</b>) Fe-doped CeO<sub>2</sub> nanoparticles, (<b>b</b>) yeast template, (<b>c</b>) CeO<sub>2</sub> hollow microspheres, and Fe-doped CeO<sub>2</sub> hollow microspheres before (<b>d</b>) and after (<b>e</b>–<b>h</b>) calcination. Figure reproduced from ref. [<a href="#B211-nanomaterials-14-01190" class="html-bibr">211</a>]. The SEM image (<b>i</b>) and TEM images (<b>j</b>–<b>l</b>) of Fe<sub>3</sub>O<sub>4</sub>@MnO<sub>2</sub> BBHs. Figure reproduced from ref. [<a href="#B212-nanomaterials-14-01190" class="html-bibr">212</a>].</p>
Full article ">
24 pages, 7321 KiB  
Article
Catalytic Ozonation of Pharmaceuticals Using CeO2-CeTiOx-Doped Crossflow Ultrafiltration Ceramic Membranes
by Nikoletta Tsiarta, Silvia Morović, Vilko Mandić, Ivana Panžić, Roko Blažic, Lidija Ćurković and Wolfgang Gernjak
Nanomaterials 2024, 14(13), 1163; https://doi.org/10.3390/nano14131163 - 7 Jul 2024
Viewed by 922
Abstract
The removal of persistent organic micropollutants (OMPs) from secondary effluent in wastewater treatment plants is critical for meeting water reuse standards. Traditional treatment methods often fail to adequately degrade these contaminants. This study explored the efficacy of a hybrid ozonation membrane filtration (HOMF) [...] Read more.
The removal of persistent organic micropollutants (OMPs) from secondary effluent in wastewater treatment plants is critical for meeting water reuse standards. Traditional treatment methods often fail to adequately degrade these contaminants. This study explored the efficacy of a hybrid ozonation membrane filtration (HOMF) process using CeO2 and CeTiOx-doped ceramic crossflow ultrafiltration ceramic membranes for the degradation of OMPs. Hollow ceramic membranes (CM) with a 300 kDa molecular weight cut-off (MWCO) were modified to serve as substrates for catalytic nanosized metal oxides in a crossflow and inside-out operational configuration. Three types of depositions were tested: a single layer of CeO2, a single layer of CeTiOx, and a combined layer of CeO2 + CeTiOx. These catalytic nanoparticles were distributed uniformly using a solution-based method supported by vacuum infiltration to ensure high-throughput deposition. The results demonstrated successful infiltration of the metal oxides, although the yield permeability and transmembrane flow varied, following this order: pristine > CeTiOx > CeO2 > CeO2 + CeTiOx. Four OMPs were examined: two easily degraded by ozone (carbamazepine and diclofenac) and two recalcitrant (ibuprofen and pCBA). The highest OMP degradation was observed in demineralized water, particularly with the CeO2 + CeTiOx modification, suggesting O3 decomposition to hydroxyl radicals. The increased resistance in the modified membranes contributed to the adsorption phenomena. The degradation efficiency decreased in secondary effluent due to competition with the organic and inorganic load, highlighting the challenges in complex water matrices. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic representation of the hybrid ozonation–membrane filtration (HOMF) process.</p>
Full article ">Figure 2
<p>Solid-state X-ray diffraction (SS-XRD) patterns of the modified ceramic membranes pristine (black), CeO<sub>2</sub> (green) CeTiO<sub>x</sub> (blue), and CeO<sub>2</sub> + CeTiO<sub>x</sub> (red).</p>
Full article ">Figure 3
<p>Surface morphologies with Scanning Electron Microscopy (SEM) of the inner surface of the membrane (<b>left</b>) and cross-section (<b>right</b>) for the three modifications (<b>a</b>,<b>d</b>) CeO<sub>2</sub>, (<b>b</b>,<b>e</b>) CeTiO<sub>x</sub>, and (<b>c</b>,<b>f</b>) CeO<sub>2</sub> + CeTiO<sub>x</sub> in various magnifications (50×, 500×, 5000×, and 10,000×).</p>
Full article ">Figure 4
<p>SEM-EDS mapping (250× magnification) of the inner surface of the modified ceramic membranes for (<b>a</b>) CeO<sub>2</sub>, (<b>b</b>) CeTiO<sub>x</sub>, (<b>c</b>) CeO<sub>2</sub> + CeTiO<sub>x</sub>, and the cross-section for (<b>d</b>) CeO<sub>2</sub>.</p>
Full article ">Figure 5
<p>Amplitude (<b>top</b>) and three-dimensional (3D) representation (<b>bottom</b>) of the surface of the unmodified and modified ceramic membranes: (<b>a</b>) pristine, (<b>b</b>) CeO<sub>2</sub>, (<b>c</b>) CeTiO<sub>x</sub>, and (<b>d</b>) CeO<sub>2</sub> + CeTiO<sub>x</sub>. (X and Y axes represent distance in μm).</p>
Full article ">Figure 6
<p>(<b>a</b>) Differential intrusion curves and (<b>b</b>) permeable pore volume for unmodified and modified ceramic membranes.</p>
Full article ">Figure 7
<p>Transmembrane flow of the ceramic membranes at 25 °C and 1 bar (L m<sup>−2</sup> h<sup>−1</sup>) before and after the deposition of the metal oxides.</p>
Full article ">Figure 8
<p>Degradation of the pharmaceuticals (<b>a</b>) CBZ, (<b>b</b>) DCF, (<b>c</b>) IBP, and (<b>d</b>) pCBA using the CM 300 kDa MWCO with different surface modifications (Operating conditions: [OMPs]<sub>in</sub> = 10 μM, [NaHCO<sub>3</sub>] = 1 mM, [TOD] = 4 mg L<sup>−1</sup>).</p>
Full article ">Figure 9
<p>Degradation of the pharmaceuticals (<b>a</b>) CBZ, (<b>b</b>) DCF, (<b>c</b>) IBP, and (<b>d</b>) pCBA using the CM 300 kDa MWCO with different surface modifications in the presence of TBA, a hydroxyl radical scavenger at 1:1 molar ratio relative to the total OMPs (Operating conditions: [OMPs]<sub>in</sub> = 10 μM, [TBA] = 0.05 mM, [NaHCO<sub>3</sub>] = 1 mM, [TOD] = 4 mg L<sup>−1</sup>).</p>
Full article ">Figure 10
<p>Degradation of the pharmaceuticals (<b>a</b>) CBZ, (<b>b</b>) DCF, (<b>c</b>) IBP, and (<b>d</b>) pCBA with ozonation alone in the presence of bicarbonate and/or scavenger (Operating conditions: [OMPs]<sub>in</sub> = 10 μM, [TBA] = 0.05 mM (1:1) or 0.5 mM (1:10), [NaHCO<sub>3</sub>] = 1 mM, [TOD] = 4 mg L<sup>−1</sup>).</p>
Full article ">Figure 11
<p>Degradation of the pharmaceuticals (<b>a</b>) CBZ, (<b>b</b>) DCF, (<b>c</b>) IBP, and (<b>d</b>) pCBA using the pristine and CeO<sub>2</sub> + CeTiO<sub>x</sub> CM 300 kDa MWCO in secondary effluent (Operating conditions: [OMPs]<sub>in</sub> = 10 μM, [TOD] = 4 mg L<sup>−1</sup>).</p>
Full article ">Figure 12
<p>Permeability recovery after adsorption and ozonation experiments with secondary effluent for pristine (black dots) and CeO<sub>2</sub> + CeTiO<sub>x</sub>-modified (red diamonds) ceramic membrane.</p>
Full article ">Figure 13
<p>Residual ozone concentration [mg L<sup>−1</sup>] in permeate flow for pristine and CeO<sub>2</sub> + CeTiO<sub>x</sub>-modified membrane with demineralized water, with TBA at 1:1 molar ratio and Girona WWTP secondary effluent (ozone residual concentration was determined with the indigo method at 600 nm).</p>
Full article ">
16 pages, 312 KiB  
Review
Application of Engineered Nanomaterials as Nanocatalysts in Catalytic Ozonation: A Review
by Rita M. F. Cardoso, Joaquim C. G. Esteves da Silva and Luís Pinto da Silva
Materials 2024, 17(13), 3185; https://doi.org/10.3390/ma17133185 - 28 Jun 2024
Viewed by 780
Abstract
Given the growing scarcity of water and the continuous increase in emerging pollutants detected in water bodies, there is an imperative need to develop new, more effective, and sustainable treatments for wastewater. Advanced oxidation processes (AOPs) are considered a competitive technology for water [...] Read more.
Given the growing scarcity of water and the continuous increase in emerging pollutants detected in water bodies, there is an imperative need to develop new, more effective, and sustainable treatments for wastewater. Advanced oxidation processes (AOPs) are considered a competitive technology for water treatment. Specifically, ozonation has received notable attention as a promising approach for degrading organic pollutants in wastewater. However, different groups of pollutants are hardly degradable via single ozonation. With continuous development, it has been shown that using engineered nanomaterials as nanocatalysts in catalytic ozonation can increase efficiency by turning this process into a low-selective AOP for pollutant degradation. Nanocatalysts promote ozone decomposition and form active free radicals responsible for increasing the degradation and mineralization of pollutants. This work reviews the performances of different nanomaterials as homogeneous and heterogeneous nanocatalysts in catalytic ozonation. This review focuses on applying metal- and carbon-based engineered nanomaterials as nanocatalysts in catalytic ozonation and on identifying the main future directions for using this type of AOP toward wastewater treatment. Full article
(This article belongs to the Special Issue Advanced Luminescent Materials and Applications)
16 pages, 20893 KiB  
Article
Degradation of Sodium Acetate by Catalytic Ozonation Coupled with MnOx/NiOOH-Modified Fly Ash
by Ruifu Chen, Hao Zhang, Shengyu Shao, Huajun Xu, Kaicheng Zhou, Yinzhi Jiang and Pengfei Sun
Toxics 2024, 12(6), 412; https://doi.org/10.3390/toxics12060412 - 4 Jun 2024
Viewed by 694
Abstract
Fly ash, a type of solid waste generated in power plants, can be utilized as a catalyst carrier to enhance its value-added potential. Common methods often involve using a large amount of alkali for preprocessing, resulting in stable quartz and mullite forming silicate [...] Read more.
Fly ash, a type of solid waste generated in power plants, can be utilized as a catalyst carrier to enhance its value-added potential. Common methods often involve using a large amount of alkali for preprocessing, resulting in stable quartz and mullite forming silicate dissolution. This leads to an increased specific surface area and pore structure. In this study, we produced a catalyst composed of MnOx/NiOOH supported on fly ash by directly employing nickel hydroxide and potassium permanganate to generate metal active sites over the fly ash surface while simultaneously creating a larger specific surface area and pore structure. The ozone catalytic oxidation performance of this catalyst was evaluated using sodium acetate as the target organic matter. The experimental results demonstrated that an optimal removal efficiency of 57.5% for sodium acetate was achieved, surpassing even that of MnOx/NiOOH supported catalyst by using γ-Al2O3. After loading of MnOx/NiOOH, an oxygen vacancy is formed on the surface of fly ash, which plays an indirect oxidation effect on sodium acetate due to the transformation of ozone to •O2 and •OH over this oxygen vacancy. The reaction process parameters, including varying concentrations of ozone, sodium acetate, and catalyst dosage, as well as pH value and the quantitative analysis of formed free radicals, were examined in detail. This work demonstrated that fly ash could be used as a viable catalytic material for wastewater treatment and provided a new solution to the added value of fly ash. Full article
(This article belongs to the Special Issue Effective Catalytic Processes for Water and Wastewater Treatment)
Show Figures

Figure 1

Figure 1
<p>Preparation process of fly ash catalyst.</p>
Full article ">Figure 2
<p>Diagram of a device for catalytic ozone oxidation reaction.</p>
Full article ">Figure 3
<p>(<b>a</b>) The degradation performance of sodium acetate over samples with different ratios of Ni/Mn; (<b>b</b>) the degradation performance of sodium acetate over samples with different loading amounts of metal species; (<b>c</b>) the degradation performance of sodium acetate over samples with different catalyst carriers; (ozone concentration: 50 mg/L, catalyst dosage: 3 g/L, sodium acetate concentration: 150 mg/L, pH: 7.5, and temperature: 25 °C).</p>
Full article ">Figure 4
<p>XRD spectra of pretreated fly ash and catalysts with different Ni/Mn ratios.</p>
Full article ">Figure 5
<p>SEM image: (<b>a1</b>–<b>a3</b>) CFA, (<b>b1</b>–<b>b3</b>) ACFA, (<b>c1</b>–<b>c3</b>) 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA; (<b>d</b>) EDS-mapping of 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA.</p>
Full article ">Figure 6
<p>HR-TEM images of catalyst 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA.</p>
Full article ">Figure 7
<p>XPS spectra of pretreated fly ash and catalysts with different Ni/Mn ratios: (<b>a</b>) survey spectrum, (<b>b</b>) Ni 2p, (<b>c</b>) Mn 2p, (<b>d</b>) O 1s.</p>
Full article ">Figure 8
<p>ESR spectra of pretreated fly ash and catalysts with different loading amounts of metal species.</p>
Full article ">Figure 9
<p>Exploring the influence of various reaction parameters on the efficiency of sodium acetate degradation using a 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA catalyst: (<b>a</b>) ozone concentration; (<b>b</b>) catalyst dosage; and (<b>c</b>) sodium acetate concentration (ozone concentration: 50 mg/L, catalyst dosage: 3 g/L, sodium acetate concentration: 150 mg/L, pH: 7.5, temperature: 25 °C).</p>
Full article ">Figure 10
<p>Effects of different pH values on the degradation performance of sodium acetate by using catalyst 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA (ozone concentration: 50 mg/L, catalyst dosage: 3 g/L, sodium acetate concentration: 150 mg/L, pH: 7.5, temperature: 25 °C).</p>
Full article ">Figure 11
<p>Effect of removal efficiency of sodium acetate over catalyst 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA with the addition of different quenchers. (ozone concentration: 50 mg/L, catalyst dosage: 3 g/L, sodium acetate concentration: 150 mg/L, pH: 7.5, temperature: 25 °C, TBA = 5 mmol/L, p-BQ = 0.5 mmol/L, L-HIS = 1 mmol/L).</p>
Full article ">Figure 12
<p>EPR spectra of 0.2-Ni<sub>0.67</sub>Mn<sub>0.33</sub>OOH/ACFA: (<b>a</b>) DMPO-•O<sub>2</sub><sup>–</sup>, (<b>b</b>) DMPO-•OH, (<b>c</b>) TEMP-<sup>1</sup>O<sub>2</sub>.</p>
Full article ">Figure 13
<p>(<b>a</b>) UV profile of NBT solution; (<b>b</b>) NBT absorption spectroscopy to determine the concentration of •O<sub>2</sub><sup>–</sup>; (<b>c</b>) fluorescence profile of TAOH solution; (<b>d</b>) Photoluminescence spectroscopy of PL to determine the concentration of •OH.</p>
Full article ">
25 pages, 5084 KiB  
Article
Molecular Mechanism of Exogenous ABA to Enhance UV-B Resistance in Rhododendron chrysanthum Pall. by Modulating Flavonoid Accumulation
by Wang Yu, Fushuai Gong, Hongwei Xu and Xiaofu Zhou
Int. J. Mol. Sci. 2024, 25(10), 5248; https://doi.org/10.3390/ijms25105248 - 11 May 2024
Cited by 2 | Viewed by 790
Abstract
With the depletion of the ozone layer, the intensity of ultraviolet B (UV-B) radiation reaching the Earth’s surface increases, which in turn causes significant stress to plants and affects all aspects of plant growth and development. The aim of this study was to [...] Read more.
With the depletion of the ozone layer, the intensity of ultraviolet B (UV-B) radiation reaching the Earth’s surface increases, which in turn causes significant stress to plants and affects all aspects of plant growth and development. The aim of this study was to investigate the mechanism of response to UV-B radiation in the endemic species of Rhododendron chrysanthum Pall. (R. chrysanthum) in the Changbai Mountains and to study how exogenous ABA regulates the response of R. chrysanthum to UV-B stress. The results of chlorophyll fluorescence images and OJIP kinetic curves showed that UV-B radiation damaged the PSII photosystem of R. chrysanthum, and exogenous ABA could alleviate this damage to some extent. A total of 2148 metabolites were detected by metabolomics, of which flavonoids accounted for the highest number (487, or 22.67%). KEGG enrichment analysis of flavonoids that showed differential accumulation by UV-B radiation and exogenous ABA revealed that flavonoid biosynthesis and flavone and flavonol biosynthesis were significantly altered. GO analysis showed that most of the DEGs produced after UV-B radiation and exogenous ABA were distributed in the cellular process, cellular anatomical entity, and catalytic activity. Network analysis of key DFs and DEGs associated with flavonoid synthesis identified key flavonoids (isorhamnetin-3-O-gallate and dihydromyricetin) and genes (TRINITY_DN2213_c0_g1_i4-A1) that promote the resistance of R. chrysanthum to UV-B stress. In addition, multiple transcription factor families were found to be involved in the regulation of the flavonoid synthesis pathway under UV-B stress. Overall, R. chrysanthum actively responded to UV-B stress by regulating changes in flavonoids, especially flavones and flavonols, while exogenous ABA further enhanced its resistance to UV-B stress. The experimental results not only provide a new perspective for understanding the molecular mechanism of the response to UV-B stress in the R. chrysanthum, but also provide a valuable theoretical basis for future research and application in improving plant adversity tolerance. Full article
(This article belongs to the Special Issue Advance in Plant Abiotic Stress)
Show Figures

Figure 1

Figure 1
<p>Chlorophyll fluorescence images of <span class="html-italic">R. chrysanthum</span> photosystem II (PS II) under different treatments. The images (<b>A</b>,<b>C</b>,<b>E</b>) represent Fv/Fm, Y (II), and Fo, respectively. (<b>B</b>,<b>D</b>,<b>F</b>) are bar graphs representing average values for Fv/Fm, Y (II), and Fo, respectively. The height of each bar graph represents an average of three copies, and the length of each error bar represents a corresponding standard deviation. After the ANOVA test, the difference between different letters is significant (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">Figure 2
<p>Effect of UV-B radiation and exogenous ABA on OJIP curves of <span class="html-italic">R. chrysanthum</span> leaves. (<b>A</b>) OJIP raw curves; (<b>B</b>) O-P point-normalised Vt curves; (<b>C</b>) The difference ΔVt between UV-B and control PAR and ABA versus UV-B curves; (<b>D</b>) The difference ΔWt between UV-B and control PAR and ABA versus UV-B curves.</p>
Full article ">Figure 3
<p>Information related to the detected metabolites of <span class="html-italic">R. chrysanthum</span> under different treatments. (<b>A</b>) OPLS-DA (orthogonal partial least squares discriminant analysis) of MN group metabolites; (<b>B</b>) OPLS-DA of NQ group metabolites; (<b>C</b>) 2148 metabolite classes make up the ring. Each color represents a metabolite class, and the area of the color block indicates the percentage of that class; (<b>D</b>) Statistical pie chart of the detected flavonoids and their percentages.</p>
Full article ">Figure 4
<p>Information of DFs of <span class="html-italic">R. chrysanthum</span> under different treatments. (<b>A</b>) Statistics of the number of up-regulated and down-regulated DFs under UV-B and exogenous ABA treatments; (<b>B</b>) Wayne diagram of the DFs under UV-B radiation and exogenous ABA treatments; (<b>C</b>) Heatmap of clustering of the 17 common DFs; (<b>D</b>) Statistics of the number of types of up-regulated DFs; (<b>E</b>) Statistics of the number of types of down-regulated DFs; (<b>F</b>) KEGG enrichment analysis of the DFs in MvsN KEGG enrichment analysis; (<b>G</b>) KEGG enrichment analysis of DFs in NvsQ.</p>
Full article ">Figure 4 Cont.
<p>Information of DFs of <span class="html-italic">R. chrysanthum</span> under different treatments. (<b>A</b>) Statistics of the number of up-regulated and down-regulated DFs under UV-B and exogenous ABA treatments; (<b>B</b>) Wayne diagram of the DFs under UV-B radiation and exogenous ABA treatments; (<b>C</b>) Heatmap of clustering of the 17 common DFs; (<b>D</b>) Statistics of the number of types of up-regulated DFs; (<b>E</b>) Statistics of the number of types of down-regulated DFs; (<b>F</b>) KEGG enrichment analysis of the DFs in MvsN KEGG enrichment analysis; (<b>G</b>) KEGG enrichment analysis of DFs in NvsQ.</p>
Full article ">Figure 5
<p>Multi-perspective analysis of differentially expressed genes (DEGs) among different controls. (<b>A</b>) Wayne plots of DEGs between Groups M and N and between Groups N and Q; (<b>B</b>) Statistical volcano plots of the number of upward and downward adjustments of DEGs in Groups M and N; (<b>C</b>) Statistical volcano plots of the number of upward and downward adjustments of DEGs in Groups N and Q.</p>
Full article ">Figure 6
<p>Dynamic changes of flavonoid-related pathways under two treatments. (<b>A</b>) Statistics of the number of DFs and DEGs in the two flavonoid-related pathways; (<b>B</b>) Diagram of the network pathways of DFs and DEGs in the flavonoid-related pathway.</p>
Full article ">Figure 6 Cont.
<p>Dynamic changes of flavonoid-related pathways under two treatments. (<b>A</b>) Statistics of the number of DFs and DEGs in the two flavonoid-related pathways; (<b>B</b>) Diagram of the network pathways of DFs and DEGs in the flavonoid-related pathway.</p>
Full article ">Figure 7
<p>Correlation analysis of key DFs and DEGs. (<b>A</b>) Correlation network diagram between DFs and DEGs under UV-B and exogenous ABA treatments. The critical value for correlation analysis was r<sup>2</sup> value ≥ 0.8 and <span class="html-italic">p</span> value &lt; 0.05; blue triangles represent DFs, yellow circles represent DEGs, yellow lines indicate positive correlations, and blue lines indicate negative correlations. (<b>B</b>) Box plots of the expression of key DFs under UV-B and exogenous ABA treatments. Box plots of the expression of key DFs under UV-B and exogenous ABA treatments. (<b>C</b>) Statistics on the number of species of TFs that have been detected in <span class="html-italic">R. chrysanthum</span>. (<b>D</b>) Correlation analysis and chord plots of key differential TFs with enzymes involved in differential accumulation in the flavonoid pathway. Red lines indicate positive correlations and blue lines indicate negative correlations.</p>
Full article ">Figure 7 Cont.
<p>Correlation analysis of key DFs and DEGs. (<b>A</b>) Correlation network diagram between DFs and DEGs under UV-B and exogenous ABA treatments. The critical value for correlation analysis was r<sup>2</sup> value ≥ 0.8 and <span class="html-italic">p</span> value &lt; 0.05; blue triangles represent DFs, yellow circles represent DEGs, yellow lines indicate positive correlations, and blue lines indicate negative correlations. (<b>B</b>) Box plots of the expression of key DFs under UV-B and exogenous ABA treatments. Box plots of the expression of key DFs under UV-B and exogenous ABA treatments. (<b>C</b>) Statistics on the number of species of TFs that have been detected in <span class="html-italic">R. chrysanthum</span>. (<b>D</b>) Correlation analysis and chord plots of key differential TFs with enzymes involved in differential accumulation in the flavonoid pathway. Red lines indicate positive correlations and blue lines indicate negative correlations.</p>
Full article ">Figure 8
<p>Flowchart of experimental treatment of <span class="html-italic">R. chrysanthum</span>.</p>
Full article ">
19 pages, 6977 KiB  
Article
Deflamin Attenuated Lung Tissue Damage in an Ozone-Induced COPD Murine Model by Regulating MMP-9 Catalytic Activity
by Elia Ana Baltazar-García, Belinda Vargas-Guerrero, Ana Lima, Ricardo Boavida Ferreira, María Luisa Mendoza-Magaña, Mario Alberto Ramírez-Herrera, Tonatiuh Abimael Baltazar-Díaz, José Alfredo Domínguez-Rosales, Adriana María Salazar-Montes and Carmen Magdalena Gurrola-Díaz
Int. J. Mol. Sci. 2024, 25(10), 5063; https://doi.org/10.3390/ijms25105063 - 7 May 2024
Viewed by 1077
Abstract
Chronic obstructive pulmonary disease (COPD) is comprised of histopathological alterations such as pulmonary emphysema and peribronchial fibrosis. Matrix metalloproteinase 9 (MMP-9) is one of the key enzymes involved in both types of tissue remodeling during the development of lung damage. In recent studies, [...] Read more.
Chronic obstructive pulmonary disease (COPD) is comprised of histopathological alterations such as pulmonary emphysema and peribronchial fibrosis. Matrix metalloproteinase 9 (MMP-9) is one of the key enzymes involved in both types of tissue remodeling during the development of lung damage. In recent studies, it was demonstrated that deflamin, a protein component extracted from Lupinus albus, markedly inhibits the catalytic activity of MMP-9 in experimental models of colon adenocarcinoma and ulcerative colitis. Therefore, in the present study, we investigated for the first time the biological effect of deflamin in a murine COPD model induced by chronic exposure to ozone. Ozone exposure was carried out in C57BL/6 mice twice a week for six weeks for 3 h each time, and the treated group was orally administered deflamin (20 mg/kg body weight) after each ozone exposure. The histological results showed that deflamin attenuated pulmonary emphysema and peribronchial fibrosis, as evidenced by H&E and Masson’s trichrome staining. Furthermore, deflamin administration significantly decreased MMP-9 activity, as assessed by fluorogenic substrate assay and gelatin zymography. Interestingly, bioinformatic analysis reveals a plausible interaction between deflamin and MMP-9. Collectively, our findings demonstrate the therapeutic potential of deflamin in a COPD murine model, and suggest that the attenuation of the development of lung tissue damage occurs by deflamin-regulated MMP-9 catalytic activity. Full article
(This article belongs to the Special Issue Natural Compounds in Health and Disease)
Show Figures

Figure 1

Figure 1
<p>Inhibitory effect of deflamin on MMP-9 activity evaluated by fluorescent gelatin in vitro assay. <span class="html-italic">L. albus</span> deflamin was added to evaluate its inhibitory effect on the gelatinolytic activity of recombinant human MMP-9 (rMMP-9). Bars indicate the percentage of enzymatic activity relative to the control (rMMP-9 without inhibitor) and represent the averages of at least three experiments performed in triplicate. Results are expressed as means ± SEM. **** <span class="html-italic">p</span> &lt; 0.0001; * <span class="html-italic">p</span> &lt; 0.05; ns, not significant; DEFL, 98 μM deflamin; QRT, 750 μM quercetin; PHEN, 0.5 μM 1,10-phenanthroline.</p>
Full article ">Figure 2
<p>Comparison of histological lung changes in the experimental groups. (<b>a</b>) Representative photomicrographs from lung tissue stained with hematoxylin and eosin (H&amp;E) at magnifications 4× and 10× that allowed us to assess the presence and severity of emphysema and vascular alterations among the groups. (<b>b</b>) Mean linear intercept (Lm) was quantified from the captured images and the results are expressed as means ± SEM. **** <span class="html-italic">p</span> &lt; 0.0001; ns, not significant. OZN, ozone group; OZN + DEFL, ozone group treated with deflamin (20 mg/kg BW).</p>
Full article ">Figure 3
<p>Peribronchial collagen deposition in lung tissues from the three experimental groups. (<b>a</b>) Representative photomicrographs from lung tissue stained with Masson’s trichrome staining at magnifications 10× and 40× to allow identification the presence and distribution of lung tissue fibrosis among the groups, especially in the peribronchial area. Collagen deposition is identified in blue–purple. (<b>b</b>) The percentage of peribronchial fibrosis was quantified from the captured images, and the results were expressed as means ± SEM. **** <span class="html-italic">p</span> &lt; 0.0001; ns, not significant; OZN, ozone group; OZN + DEFL, ozone group treated with deflamin (20 mg/kg BW).</p>
Full article ">Figure 4
<p>Matrix metalloproteinase (MMP)-9 protein expression in mouse lung tissue homogenates analyzed by Western blotting. (<b>a</b>) The protein bands corresponding to MMP-9 and β-actin are shown. (<b>b</b>) The quantitative analysis indicates a downward trend in the protein expression of MMP-9 in the group exposed to ozone and treated with deflamin compared to the ozone group. Data are expressed as means ± SEM. * <span class="html-italic">p</span> &lt; 0.05; ns, not significant; OZN, ozone group; OZN + DEFL, ozone group treated with deflamin (20 mg/kg BW).</p>
Full article ">Figure 5
<p>Matrix metalloproteinase (MMP)-9 activity of mouse lung tissue homogenates analyzed by gelatin zymography. (<b>a</b>) Representative gelatin zymography of lung tissue homogenates in the three experimental groups. (<b>b</b>) MMP-9 activity comparison (density arbitrary units) shows that deflamin statistically reduced MMP-9 activity in ozone-challenged mice. Data are presented as means ± SEM. * <span class="html-italic">p</span> &lt; 0.05; ns, not significant; OZN, ozone group; OZN + DEFL, ozone group treated with deflamin (20 mg/kg BW); rMMP-9, human recombinant MMP-9.</p>
Full article ">Figure 6
<p>Three-dimensional molecular interaction between MMP-9 and the δ-conglutin large subunit. (<b>a</b>) Overview of the predicted interaction between MMP-9 (orange) and the δ-conglutin large subunit (green). (<b>b</b>) Close-up of the main interaction (no hydrogen bond) between TYR160 of MMP-9 and ASN71 of the δ-conglutin large subunit. (<b>c</b>) Close-up of the four main hydrogen bonds between MMP-9 and the δ-conglutin large subunit. (<b>d</b>) Close-up of two of the main hydrogen bonds in which ASN71 of the δ-conglutin large subunit was involved. Proteins are represented in their secondary structures. Calcium ions are in green and zinc ions in gray. Both the MMP-9 active site and the residues that are interacting, between proteins, are represented in sticks. Hydrogen bond distances are presented in Å and as black lines.</p>
Full article ">
18 pages, 11384 KiB  
Article
Preparation of CeO2 Supported on Graphite Catalyst and Its Catalytic Performance for Diethyl Phthalate Degradation during Ozonation
by Xin-Yi Tao, Yu-Hong Cui and Zheng-Qian Liu
Water 2024, 16(9), 1274; https://doi.org/10.3390/w16091274 - 29 Apr 2024
Viewed by 824
Abstract
Catalysts for the efficient catalytic decomposition of ozone to generate reactive free radicals to oxidize pollutants are needed. The graphite-supported CeO2 catalyst was optimally prepared, and its activity in ozonation was evaluated using the degradation of diethyl phthalate (DEP) as an index. [...] Read more.
Catalysts for the efficient catalytic decomposition of ozone to generate reactive free radicals to oxidize pollutants are needed. The graphite-supported CeO2 catalyst was optimally prepared, and its activity in ozonation was evaluated using the degradation of diethyl phthalate (DEP) as an index. The stability of CeO2/graphite catalyst and the influence of operating conditions on its catalytic activity were investigated, and the mechanism of CeO2/graphite catalytic ozonation was analyzed. CeO2/graphite had the highest catalytic activity at a Ce load of 3.5% and a pyrolysis temperature of 400 °C with the DEP degradation efficiency of 75.0% and the total organic carbon (TOC) removal efficiency of 48.3%. No dissolution of active components was found during the repeated use of CeO2/graphite catalyst. The ozone dosage, catalyst dosage, initial pH, and reaction temperature have positive effects on the DEP degradation by CeO2/graphite catalytic ozonation. The presence of tert-butanol significantly inhibits the degradation of DEP at an initial pH of 3.0, 5.8, or 9.0, and the experimental results of the OH probe compound pCBA indicate that the CeO2/graphite catalyst can efficiently convert ozone into OH in solution. The DEP degradation in the CeO2/graphite catalytic ozonation mainly depends on the OH in the bulk solution formed by ozone decomposition. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of DEP degradation during ozonation with graphite, ZnO/graphite, and CeO<sub>2</sub>/graphite catalysts. (<b>a</b>) DEP concentration variation, (<b>b</b>) TOC removal, and metal ion leaching after 10 min reaction. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 2
<p>Reuse of the CeO<sub>2</sub>/graphite catalyst during ozonation for the DEP degradation and TOC removal. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 3
<p>Effect of (<b>a</b>) cerium loading amount and (<b>b</b>) pyrolysis temperature on ozonation of DEP with CeO<sub>2</sub>/graphite catalysts. Reaction conditions for ozonation: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>. Reaction conditions for catalyst preparation: T = 400 °C for (<b>a</b>), cerium loading amount = 3.5% for (<b>b</b>).</p>
Full article ">Figure 4
<p>Characterization of CeO<sub>2</sub>/graphite catalysts. (<b>a</b>) SEM (×80,000), (<b>b</b>) TEM (×100,000), XRD patterns with different (<b>c</b>) cerium loading amount and (<b>d</b>) pyrolysis temperature. Reaction conditions for catalyst preparation: T = 400 °C for (<b>a</b>–<b>c</b>), cerium loading amount = 3.5% for (<b>a</b>,<b>b</b>,<b>d</b>).</p>
Full article ">Figure 5
<p>Effect of ozone dosage on DEP degradation during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 6
<p>Effect of catalyst dosage on DEP degradation during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>.</p>
Full article ">Figure 7
<p>Effect of initial DEP concentration on DEP degradation during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial pH = 5.8, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 8
<p>Effect of initial pH on DEP degradation during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 9
<p>Effect of reaction temperature on DEP degradation during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Evolution of ozone concentration during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 11
<p>Evolution of H<sub>2</sub>O<sub>2</sub> concentration during CeO<sub>2</sub>/graphite catalytic ozonation. Reaction conditions: T = 20 °C, initial pH = 5.8, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 12
<p>Effect of TBA on ozonation of DEP with or without CeO<sub>2</sub>/graphite catalyst under different initial pH. Reaction conditions: T = 20 °C, initial DEP concentration = 3 μM, ozone gas concentration = 0.38 mg min<sup>−1</sup>, catalyst dosage = 100 mg L<sup>−1</sup>.</p>
Full article ">Figure 13
<p><span class="html-italic">p</span>CBA degradation and ozone decomposition in ozonation alone and ozonation with different catalysts. Reaction conditions: T = 20 °C, initial pH = 7.1, initial <span class="html-italic">p</span>CBA concentration = 1 μM, initial TBA concentration = 80 μM, initial ozone concentration = 2 mg L<sup>−1</sup>, catalyst dosage = 50 mg L<sup>−1</sup>.</p>
Full article ">
Back to TopTop