[go: up one dir, main page]

Next Issue
Volume 13, August
Previous Issue
Volume 13, June
 
 

Catalysts, Volume 13, Issue 7 (July 2023) – 120 articles

Cover Story (view full-size image): Cr(VI) is a common heavy metal pollutant present in the aquatic environment, which possesses toxic and carcinogenic properties. In this study, a solvothermal reaction was used to prepare porphyrin (TCPP)-modified UiO-66-NH2 (UNT). The UNT integrated adsorption and photocatalytics in the application for dealing with Cr(VI). We found that the TCPP doping amount of 15 mg UNT (15-UNT) had a 10 times higher reduction rate of Cr(VI) than pristine UiO-66-NH2. In addition, the introduction of TCPP not only enhanced the absorption of light but also enabled the transport of photogenerated electrons from TCPP to UiO-66-NH2, which promoted the separation of charge carriers. Overall, this work presented a possible relationship between the crystal structures and the performance of UNT. 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, 3859 KiB  
Article
Photothermal-Assisted Photocatalytic Degradation of Tetracycline in Seawater Based on the Black g-C3N4 Nanosheets with Cyano Group Defects
by Loic Jiresse Nguetsa Kuate, Zhouze Chen, Jialin Lu, Huabing Wen, Feng Guo and Weilong Shi
Catalysts 2023, 13(7), 1147; https://doi.org/10.3390/catal13071147 - 24 Jul 2023
Cited by 42 | Viewed by 2172
Abstract
As a broad-spectrum antibiotic, tetracycline (TC) has been continually detected in soil and seawater environments, which poses a great threat to the ecological environment and human health. Herein, a black graphitic carbon nitride (CN-B) photocatalyst was synthesized by the one-step calcination method of [...] Read more.
As a broad-spectrum antibiotic, tetracycline (TC) has been continually detected in soil and seawater environments, which poses a great threat to the ecological environment and human health. Herein, a black graphitic carbon nitride (CN-B) photocatalyst was synthesized by the one-step calcination method of urea and phloxine B for the degradation of tetracycline TC in seawater under visible light irradiation. The experimental results showed that the photocatalytic degradation rate of optimal CN-B-0.1 for TC degradation was 92% at room temperature within 2 h, which was 1.3 times that of pure CN (69%). This excellent photocatalytic degradation performance stems from the following factors: (i) ultrathin nanosheet thickness reduces the charge transfer distance; (ii) the cyanogen defect promotes photogenerated carriers’ separation; (iii) and the photothermal effect of CN-B increases the reaction temperature and enhances the photocatalytic activity. This study provides new insight into the design of photocatalysts for the photothermal-assisted photocatalytic degradation of antibiotic pollutants. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>,<b>b</b>) SEM images of CN and CN-B-0.1. (<b>c</b>,<b>d</b>) TEM images of CN and CN-B-0.1. (<b>e</b>) Elemental mapping images of CN-B-0.1.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns and (<b>b</b>) FT-IR spectra of CN and CN-B photocatalysts.</p>
Full article ">Figure 3
<p>(<b>a</b>) XPS survey spectra and high-resolution XPS spectra of (<b>b</b>) C 1s, (<b>c</b>) N 1s, and (<b>d</b>) O 1s for CN and CN-B-0.1 photocatalysts.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–vis DRS of as-prepared photocatalysts. (<b>b</b>) Measured band gap values of pure CN and CN-B samples. (<b>c</b>) Mott–Schottky plots. (<b>d</b>) Energy diagrams of CN and CN-B-0.1 photocatalysts.</p>
Full article ">Figure 5
<p>(<b>a</b>) Photocatalytic TC degradation activity of as-prepared samples in simulated seawater. (<b>b</b>,<b>c</b>) The pseudo-first-order degradation kinetic curves and corresponding degradation rate constants of as-prepared samples. (<b>d</b>) Four cycles of experiments for photocatalysis degradation of TC over CN-B-0.1 photocatalyst.</p>
Full article ">Figure 6
<p>(<b>a</b>) Photothermal infrared thermal images and (<b>b</b>) corresponding temperature curves of CN and CN-B-0.1 in simulated seawater. (<b>c</b>) Temperature and photothermal conversion efficiency of CN-B-0.1 photocatalyst. (<b>d</b>) Photocatalytic TC degradation activity of CN and CN-B-0.1 in simulated seawater at different temperatures. (<b>e</b>,<b>f</b>) The corresponding pseudo-first-order degradation kinetic curves and kinetic constants. (<b>g</b>) PL spectra. (<b>h</b>) Transient photocurrent response curves. (<b>i</b>) EIS plots of pristine CN and CN-B-0.1 at RT and 40 °C.</p>
Full article ">Figure 7
<p>(<b>a</b>) Degradation curves and (<b>b</b>) corresponding degradation rate of CN-B-0.1 as the photocatalyst by adding different free radical trapping agents.</p>
Full article ">Figure 8
<p>Possible mechanism of photothermal-assisted photocatalytic degradation TC in seawater by CN-B photocatalyst under visible light irradiation.</p>
Full article ">Scheme 1
<p>Schematic diagram of the synthesis process of the pure CN and CN-B nanosheets.</p>
Full article ">
18 pages, 2124 KiB  
Review
Recent Advances in Platinum and Palladium Solvent Extraction from Real Leaching Solutions of Spent Catalysts
by Ana Paula Paiva
Catalysts 2023, 13(7), 1146; https://doi.org/10.3390/catal13071146 - 24 Jul 2023
Cited by 4 | Viewed by 3655
Abstract
The strategic importance of platinum and palladium, two platinum-group metals (PGMs), is particularly supported by their technological applications, one of the most relevant being the role they perform as catalysts for several sorts of chemical reactions. The cumulative demand for these two PGMs [...] Read more.
The strategic importance of platinum and palladium, two platinum-group metals (PGMs), is particularly supported by their technological applications, one of the most relevant being the role they perform as catalysts for several sorts of chemical reactions. The cumulative demand for these two PGMs to be used as catalysts more than justifies increasing research efforts to develop sustainable recycling processes to maintain their supply. This critically appraised topic review describes the recent research trends (since 2010) developed by the world’s research communities to reach sustainable methods to recover platinum and palladium from spent catalysts in the liquid phase, namely those involving a solvent extraction (SX) step. The selected recycling processes are based on extensive fundamental research, but this paper intends to focus on information collected about SX procedures applied to real leaching samples of spent catalysts, either from automobile or industrial sources. A critical appraisal of the claimed success levels, the identified constraints, and open challenges is carried out, together with some perspectives on possible ways to redirect research efforts and minimize the gap between academia and industry on this matter. Full article
(This article belongs to the Special Issue Recent Advances Utilized in the Recycling of Catalysts II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Main applications of (<b>a</b>) platinum and (<b>b</b>) palladium (gross demand, data from 2021 [<a href="#B1-catalysts-13-01146" class="html-bibr">1</a>]).</p>
Full article ">Figure 2
<p>Solvent extraction (SX) in hydrometallurgy.</p>
Full article ">Figure 3
<p>Compounds recently developed by the Yamada group for the SX of Pd(II) and Pt(IV). Examples of structures: (<b>a</b>) acyclic thioamide derivative [<a href="#B48-catalysts-13-01146" class="html-bibr">48</a>]; (<b>b</b>) SCS pincer ligand with 2 thioamide groups [<a href="#B49-catalysts-13-01146" class="html-bibr">49</a>]; (<b>c</b>) SCS pincer ligand with 2 sulphide groups [<a href="#B53-catalysts-13-01146" class="html-bibr">53</a>]; (<b>d</b>) thiodiphenol derivative with 2-substituted amino groups [<a href="#B55-catalysts-13-01146" class="html-bibr">55</a>]; (<b>e</b>) thiophosphate-based extractant [<a href="#B60-catalysts-13-01146" class="html-bibr">60</a>]; (<b>f</b>) dithiophenol-based extractant with 3 sulphur atoms [<a href="#B65-catalysts-13-01146" class="html-bibr">65</a>].</p>
Full article ">Figure 4
<p>The structures of the thioamide (<b>a</b>) and thiodiglycolamide (<b>b</b>) derivatives investigated in [<a href="#B54-catalysts-13-01146" class="html-bibr">54</a>,<a href="#B62-catalysts-13-01146" class="html-bibr">62</a>].</p>
Full article ">Figure 5
<p>Ionic liquids investigated for Pd(II), Pt(IV), and Rh(III) extraction from leachates of spent catalysts. (<b>a</b>) Cyphos IL 101 [<a href="#B61-catalysts-13-01146" class="html-bibr">61</a>,<a href="#B63-catalysts-13-01146" class="html-bibr">63</a>,<a href="#B64-catalysts-13-01146" class="html-bibr">64</a>]; (<b>b</b>) trioctyldodecylphosphonium chloride [<a href="#B58-catalysts-13-01146" class="html-bibr">58</a>]; (<b>c</b>) trihexyl(tetradecyl)phosphonium bis(2,4,4-trimethylpentyl)phosphinate (Cyphos IL 104) [<a href="#B63-catalysts-13-01146" class="html-bibr">63</a>].</p>
Full article ">Figure 6
<p>Simplified scheme showing the integrated solvometallurgical process to recover PGMs from SACs, developed under the PLATIRUS EU Horizon 2020 project (adapted from [<a href="#B81-catalysts-13-01146" class="html-bibr">81</a>]).</p>
Full article ">
13 pages, 3674 KiB  
Article
High Performance of Nanostructured Cu2O-Based Photodetectors Grown on a Ti/Mo Metallic Substrate
by Alhoda Abdelmoneim, Mohamed Sh. Abdel-wahab, June Key Lee, Meera Moydeen Abdul Hameed, Badr M. Thamer, Abdullah M. Al-Enizi, Rayana Ibrahim Alkhalifah and Wael Z. Tawfik
Catalysts 2023, 13(7), 1145; https://doi.org/10.3390/catal13071145 - 24 Jul 2023
Cited by 1 | Viewed by 1488
Abstract
In this work, cuprous oxide (Cu2O) thin films were prepared using a simplistic sputtering technique. The films were grown on both traditional fluorine-doped tin oxide (FTO) and Ti-metallic substrates. X-ray diffraction applied for investigation of the crystal structure proved that the [...] Read more.
In this work, cuprous oxide (Cu2O) thin films were prepared using a simplistic sputtering technique. The films were grown on both traditional fluorine-doped tin oxide (FTO) and Ti-metallic substrates. X-ray diffraction applied for investigation of the crystal structure proved that the Cu2O layer acquires the cubic structure with a (111) main peak at 2θ of 36.46°. The optical absorption and transmission were detected through the utilization of a UV-Vis spectrophotometer, and the optical bandgap for the Cu2O layer was determined to be ~2.15 eV using Tauc’s equation. XPS and scanning electron microscopy were also performed for chemical structure and morphological investigation, respectively. The optoelectronic behaviors for the prepared samples were carried out using a Keithley source meter; the photocurrent density was measured in a range of applied voltage between −1 and 1 volt under the illumination of a xenon lamp with a power density of 100 mWcm−2. External quantum efficiency, sensitivity, responsivity, and detectivity were computed using proprietary models based on the experimental data. Full article
(This article belongs to the Topic Photosensitive and Optical Materials)
Show Figures

Figure 1

Figure 1
<p>The XRD diffraction patterns belong to Cu<sub>2</sub>O layer deposited on (<b>a</b>) traditional FTO and (<b>b</b>) Ti-metallic substrate.</p>
Full article ">Figure 2
<p>(<b>a</b>) The optical transmission and absorption of the Cu<sub>2</sub>O layer. (<b>b</b>) The allowed band gap (E<sub>g</sub>) estimated using Tauc’s relation.</p>
Full article ">Figure 3
<p>(<b>a</b>) The deconvoluted XPS spectra for oxygen. (<b>b</b>) The high-resolution spectrum attributed to Cu2p.</p>
Full article ">Figure 4
<p>The morphological structure for Cu<sub>2</sub>O grown on (<b>a</b>) traditional FTO and (<b>b</b>) Ti-metallic substrate.</p>
Full article ">Figure 5
<p>Optoelectronic characteristics of Cu<sub>2</sub>O/FTO photodetectors. (<b>a</b>) The photocurrent density at different wavelengths from 410 to 636 nm. (<b>b</b>) The maximum photocurrent density at every wavelength. (<b>c</b>) Responsivity. (<b>d</b>) Detectivity.</p>
Full article ">Figure 6
<p>Optoelectronic characteristics of Cu<sub>2</sub>O/Ti-Mo photodetectors. (<b>a</b>) The photocurrent density at different wavelengths from 410 to 636 nm. (<b>b</b>) The maximum photocurrent density at every wavelength. (<b>c</b>) Responsivity. (<b>d</b>) Detectivity.</p>
Full article ">Figure 7
<p>(<b>a</b>) The samples’ response to the incident light at the wavelength of 588 nm. (<b>b</b>) Dark current for both Cu<sub>2</sub>O/FTO and Cu<sub>2</sub>O/Ti samples at an applied voltage from −1 to 1 V. (<b>c</b>) the external quantum efficiency (EQE) at different wavelengths. (<b>d</b>) The variation of the photosensitivity with changing the wavelengths from 410 nm to 636 nm.</p>
Full article ">Figure 8
<p>Schematic diagram depicts the structure of the band energy of the photodetectors based on Cu<sub>2</sub>O, demonstrating how electron-hole pairs transfer during the illumination process.</p>
Full article ">Figure 9
<p>Schematic diagram of Cu<sub>2</sub>O/Ti/Mo-glass photodetector under light exposure.</p>
Full article ">
16 pages, 2685 KiB  
Article
Reaction Pathways of Gamma-Valerolactone Hydroconversion over Co/SiO2 Catalyst
by Gyula Novodárszki, Ferenc Lónyi, Magdolna R. Mihályi, Anna Vikár, Róbert Barthos, Blanka Szabó, József Valyon and Hanna E. Solt
Catalysts 2023, 13(7), 1144; https://doi.org/10.3390/catal13071144 - 23 Jul 2023
Cited by 1 | Viewed by 1508
Abstract
The hydroconversion of γ-valerolactone (GVL) over Co/SiO2 catalyst proceeds in a complex reaction network, resulting in 2-methyltetrahydrofuran (2-MTHF) as the main product, and C4–C5 alcohol and alkane side-products. The catalyst was shown to contain Co0 sites and Lewis [...] Read more.
The hydroconversion of γ-valerolactone (GVL) over Co/SiO2 catalyst proceeds in a complex reaction network, resulting in 2-methyltetrahydrofuran (2-MTHF) as the main product, and C4–C5 alcohol and alkane side-products. The catalyst was shown to contain Co0 sites and Lewis acid (Co2+ ion)/Lewis base (O2− ion) pair sites, active for hydrogenation/dehydrogenation and dehydration reactions, respectively. The initial reaction step was confirmed to be the hydrogenation of GVL to key intermediate 1,4-pentanediol (1,4-PD). Cyclodehydration of 1,4-PD led to the main product 2-MTHF, whereas its dehydration/hydrogenation gave 1-pentanol and 2-pentanol side-products, with about the same yield. In contrast, 2-pentanol was the favored alcohol product of 2-MTHF hydrogenolysis. 2-Butanol was formed by decarbonylation of 4-hydroxypentanal intermediate. The latter was the product of 1,4-PD dehydrogenation. Alkanes were formed from the alcohol side-products via dehydration/hydrogenation reactions. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRPD patterns of Co/SiO<sub>2</sub> catalyst: (a) air-calcined at 500 °C and (b) reduced in situ at 450 °C in H<sub>2</sub> flow for 1 h. Diffractograms were recorded at room temperature.</p>
Full article ">Figure 2
<p>H<sub>2</sub>-TPR curves of the Co/SiO<sub>2</sub> catalyst. The sample was either (a) pretreated in O<sub>2</sub> flow at 500 °C for 1 h, cooled to 40 °C, and flushed with N<sub>2</sub> or (b) pre-reduced in a flow of 9.0 vol% H<sub>2</sub>/N<sub>2</sub> mixture at 450 °C for 1 h, then cooled to 40 °C in the same gas mixture before the TPR run. The curves were obtained by heating up the sample at a rate of 10 °C·min<sup>−1</sup> up to 800 °C in a flow of 9.0 vol% H<sub>2</sub>/N<sub>2</sub> mixture.</p>
Full article ">Figure 3
<p>FT-IR spectra of CO adsorbed on Co/SiO<sub>2</sub> reduced in a H<sub>2</sub> flow at 450 °C in situ in the IR cell for 1 h. After reduction, the catalyst was degassed by evacuation, cooled to room temperature, and contacted with CO gas at 5 mbar pressure for 10 min. The CO gas and weakly adsorbed CO was removed by evacuation for 0.5 h, then a sample spectrum was recorded. The thin lines under the curves give the component bands obtained using a peak fitting computer program.</p>
Full article ">Figure 4
<p>FT-IR spectra of pyridine (Py) adsorbed on the (a1–a4) Co/SiO<sub>2</sub> catalyst pre-reduced in H<sub>2</sub> flow at 450 °C, and on the (b1,b2) pure silica support activated in high vacuum at 450 °C. Pre-treated samples were contacted with 5 mbar of Py vapor at 200 °C for 30 min, cooled to 100 °C and evacuated at the same temperature for 30 min. Evacuation was repeated at temperatures increasing in 100 °C intervals. Spectra were recorded at room temperature. Label L indicates the characteristic bands for Py bonded to Lewis acid sites.</p>
Full article ">Figure 5
<p>GVL conversion and product yields over Co/SiO<sub>2</sub> as a function of reaction temperature at 1.0 g<sub>cat.</sub>·g<sub>GVL</sub><sup>−1</sup>·h spacetime and 30 bar total pressure. The H<sub>2</sub>/GVL molar ratio was 12. Part (<b>A′</b>) shows the lower section of (<b>A</b>) enlarged. (2-MTHF: 2-methyltetrahydrofuran).</p>
Full article ">Figure 6
<p>GVL conversion and product yields over Co/SiO<sub>2</sub> as a function of spacetime at 200 °C (<b>A</b>,<b>A′</b>), 225 °C (<b>B</b>,<b>B′</b>) and 250 °C (<b>C</b>,<b>C′</b>) at a total pressure of 30 bar. The H<sub>2</sub>/GVL molar ratio was 12. Parts (<b>A′</b>–<b>C′</b>) show lower sections of (<b>A</b>–<b>C</b>) enlarged. (2-MTHF: 2-methyltetrahydrofuran).</p>
Full article ">Figure 7
<p>1,4-PD conversion and product yields over Co/SiO<sub>2</sub> as a function of spacetime at 200 °C (<b>A</b>,<b>A′</b>) and 225 °C (<b>B</b>,<b>B′</b>) at a total pressure of 30 bar. The H<sub>2</sub>/GVL molar ratio was 12. Sections (<b>A′</b>,<b>B′</b>) show the lower part of (<b>A</b>,<b>B</b>) enlarged. (2-MTHF: 2-methyltetrahydrofuran).</p>
Full article ">Figure 8
<p>2-MTHF conversion and product yields over Co/SiO<sub>2</sub> as a function of spacetime at 200 °C (<b>A</b>) and 225 °C (<b>B</b>) at a total pressure of 30 bar. The H<sub>2</sub>/2-MTHF molar ratio was 12. (2-MTHF: 2-methyltetrahydrofuran).</p>
Full article ">Scheme 1
<p>Activation and cyclodehydration of 1,4-PD to 2-MTHF over Lewis acid (Co<sup>2+</sup> ion)/Lewis base (O<sup>2−</sup> ion) pair sites of Co/SiO<sub>2</sub>.</p>
Full article ">Scheme 2
<p>Activation and dehydration of 1,4-PD to 4-penten-2-ol (<b>A</b>) and 3-penten-1-ol (<b>B</b>) over Lewis acid (Co<sup>2+</sup> ion)/Lewis base (O<sup>2−</sup> ion) pair sites of Co/SiO<sub>2</sub>.</p>
Full article ">Scheme 3
<p>Proposed pathways for the hydroconversion of GVL over Co/SiO<sub>2</sub> catalyst, including formation of 2-MTHF main product and C<sub>4</sub>–C<sub>5</sub> alcohol and alkane side products.</p>
Full article ">
14 pages, 2566 KiB  
Article
Polyol Synthesis of Ag-Doped Copper Oxide Nanoparticles as a Methylene Blue-Degrading Agent
by Yogeshwar Baste, Vikram Jadhav, Arpita Roy, Saad Alghamdi, Mohamed Abbas, Jari S. Algethami, Mazen Almehmadi, Mamdouh Allahyani, Devvret Verma, Krishna Kumar Yadav, Byong-Hun Jeon and Hyun-Kyung Park
Catalysts 2023, 13(7), 1143; https://doi.org/10.3390/catal13071143 - 23 Jul 2023
Cited by 9 | Viewed by 2092
Abstract
The use of metal oxide nanomaterials as photocatalysts for wastewater treatment has received significant attention in recent years due to their unique physicochemical properties. In this study, we use a polyol-mediated refluxing method to synthesize silver-incorporated copper oxide nanomaterials (Ag@CuO NMs). The use [...] Read more.
The use of metal oxide nanomaterials as photocatalysts for wastewater treatment has received significant attention in recent years due to their unique physicochemical properties. In this study, we use a polyol-mediated refluxing method to synthesize silver-incorporated copper oxide nanomaterials (Ag@CuO NMs). The use of tetra butyl ammonium bromide (TBAB) as a capping agent and ethylene glycol as a reducing agent for Ag+ to Ag is elaborated upon. The prepared Ag@CuO NMs were tested for their ability to degrade water pollutants, specifically methylene blue (MB) dye. Two different Ag contents, weights of 3% and 5%, were used to produce modified CuO-based nanomaterials. The crystalline structures of the NMs were characterized via XRD diffraction, and the morphology of the materials was investigated using FE-SEM. The optical properties were studied using UV-vis spectroscopy. The photocatalytic activity of the Ag@CuO NMs was evaluated by analyzing the degradation of MB dye when exposed to UV-visible light. Our results showed that the 5% weight Ag@CuO NM sample exhibited the most efficient degradation activity against MB dye. Therefore, these nanomaterials hold potential for photocatalytic applications, particularly for wastewater purification. Full article
Show Figures

Figure 1

Figure 1
<p>UV-Vis spectra of polyol-mediated Ag@CuO NMs.</p>
Full article ">Figure 2
<p>FTIR spectrum of polyol-mediated compound before calcination.</p>
Full article ">Figure 3
<p>FTIR spectrum of polyol-mediated Ag@CuO NPs (after calcination).</p>
Full article ">Figure 4
<p>XRD pattern of polyol-mediated Ag@CuO NPs.</p>
Full article ">Figure 5
<p>FE-SEM images of polyol-mediated Ag@CuO NPs (<b>a</b>) before calcination and (<b>b</b>) after calcination.</p>
Full article ">Figure 6
<p>Photocatalytic dye degradation mechanism using Ag@CuO NPs.</p>
Full article ">Figure 7
<p>Absorbance vs. wavelength plot. (<b>A</b>) Dye degradation using material before calcination and (<b>B</b>) dye degradation using polyol-mediated Ag@CuO NPs (after calcination).</p>
Full article ">Figure 8
<p>Absorbance vs. time of MB dye degradation.</p>
Full article ">Figure 9
<p>Synthesis of polyol-mediated Ag@CuO NPs.</p>
Full article ">
11 pages, 4665 KiB  
Article
Improving Separation Efficiency of Photogenerated Charges through Combination of Conductive Polythiophene for Selective Production of CH4
by Yiqiang Deng, Lingxiao Tu, Ping Wang, Shijian Chen, Man Zhang, Yong Xu and Weili Dai
Catalysts 2023, 13(7), 1142; https://doi.org/10.3390/catal13071142 - 23 Jul 2023
Cited by 2 | Viewed by 1153
Abstract
In today’s society, mankind is confronted with two major problems: the energy crisis and the greenhouse effect. Artificial photosynthesis can use solar energy to convert the greenhouse gas CO2 into high-value compounds, which is an ideal solution to alleviate the energy crisis [...] Read more.
In today’s society, mankind is confronted with two major problems: the energy crisis and the greenhouse effect. Artificial photosynthesis can use solar energy to convert the greenhouse gas CO2 into high-value compounds, which is an ideal solution to alleviate the energy crisis and solve the problem of global warming. The combination of ZnO and polythiophenes (PTh) can make up for each other’s drawbacks, thus improving the photoresponse behavior and separation efficiency of the photogenerated carriers. The PTh layer can transfer photogenerated electrons to ZnO, thereby extending the lifetime of the photogenerated charges. The production rate of CH4 from the photoreduction of CO2 with ZnO/PTh10 is 4.3 times that of pure ZnO, and the selectivity of CH4 is increased from 70.2% to 92.2%. The conductive PTh can absorb photons to induce π–π* transition, and the photogenerated electrons can transfer from the LUMO to the conduction band (CB) of ZnO, thus more electrons are involved in the reduction of CO2. Full article
(This article belongs to the Special Issue Photocatalysis in Air Purification)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic diagram of preparation of ZnO/PTh<sub>x</sub> (<b>a</b>). XRD patterns of ZnO and ZnO/PTh<sub>10</sub> (<b>b</b>). TEM image of ZnO/PTh<sub>10</sub> (<b>c</b>) and HRTEM image of ZnO/PTh<sub>10</sub> (<b>d</b>). HAADF-STEM of ZnO/PTh<sub>10</sub> (<b>e</b>) and corresponding elemental mapping images (<b>f</b>–<b>i</b>).</p>
Full article ">Figure 2
<p>N<sub>2</sub> adsorption–desorption isotherm diagrams of ZnO and ZnO/PTh<sub>10</sub> (<b>a</b>) and corresponding pore size distribution (<b>b</b>). Ultraviolet–visible DRS of ZnO and ZnO/PTh<sub>10</sub> (<b>c</b>) and corresponding plots of (αhν)<sup>2</sup> versus energy (hν) for the band gap energy (<b>d</b>). Valence band XPS spectra (<b>e</b>). Band structure alignments of ZnO and ZnO/PTh<sub>10</sub> (<b>f</b>).</p>
Full article ">Figure 3
<p>High-resolution XPS spectra of (<b>a</b>) Zn 2p and (<b>b</b>) O 1s. Steady-state PL spectra (<b>c</b>) and time-resolved PL decay spectra (<b>d</b>) for ZnO and ZnO/PTh<sub>10</sub>. (<b>e</b>) Transient photocurrent spectra. (<b>f</b>) Transient photocurrent spectra with ZnO/PTh<sub>10</sub> as catalyst under the atmosphere of Ar and CO<sub>2</sub>.</p>
Full article ">Figure 4
<p>The effect of different PTh loadings on the photocatalytic activity of ZnO (<b>a</b>); the illumination time is 4 h. The amount of CH<sub>4</sub> (<b>b</b>) and CO (<b>c</b>) produced by photocatalytic reduction of CO<sub>2</sub> over time. (<b>d</b>) Mass spectrometry analysis of reduction products.</p>
Full article ">Figure 5
<p>(<b>a</b>) Mott–Schottky plots of ZnO and ZnO/PTh<sub>10</sub>. (<b>b</b>) Speculated mechanism for photocatalytic reduction of CO<sub>2</sub> over ZnO/PTh<sub>10</sub> under visible light irradiation.</p>
Full article ">
19 pages, 7460 KiB  
Article
A Comparative Study of the Antiviral Properties of Thermally Sprayed Coatings against Human Coronavirus HCoV-229E
by Elnaz Alebrahim, Hediyeh Khatibnezhad, Morvarid Mohammadian Bajgiran, Magan Solomon, Chen Liang, Selena M. Sagan, Rogerio S. Lima, Jörg Oberste Berghaus, Maniya Aghasibeig and Christian Moreau
Catalysts 2023, 13(7), 1141; https://doi.org/10.3390/catal13071141 - 22 Jul 2023
Cited by 3 | Viewed by 1490
Abstract
For decades, novel viral strains of respiratory tract infections have caused human pandemics and initiated widespread illnesses. The recent coronavirus disease 2019 (COVID-19) outbreak caused by the SARS-CoV-2 virus has raised an urgent need to develop novel antiviral coatings as one of the [...] Read more.
For decades, novel viral strains of respiratory tract infections have caused human pandemics and initiated widespread illnesses. The recent coronavirus disease 2019 (COVID-19) outbreak caused by the SARS-CoV-2 virus has raised an urgent need to develop novel antiviral coatings as one of the potential solutions to mitigate the transmission of viral pathogens. Titanium dioxide is considered an excellent candidate for viral disinfection under light irradiation, with the potential to be activated under visible light for indoor applications. This research assessed the antiviral performance of thermally sprayed TiO2 coatings under UVA and ambient light. We also report the antiviral performance of TiO2 composites with other oxides, such as Cu2O and Al2O3, produced by suspension plasma spray, atmospheric plasma spray, and suspension high-velocity oxygen fuel techniques. To evaluate the antiviral performance of the above coatings in a containment level-2 laboratory, a human common cold coronavirus, HCoV-229E, was initially used as a relevant surrogate for SARS-CoV-2. Coatings were also analyzed using SEM and XRD and were classified based on their surface roughness, porosity, and phase composition. Collectively, the thermally sprayed coatings showed comparable or slightly better antiviral activity compared to copper. The most significant level of activity observed was approximately 20% to 50% higher than that of a pure copper plate. Full article
(This article belongs to the Special Issue Photocatalytic Nanomaterials for Abatement of Microorganisms)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of coatings as (a) T2-APS, (b) T3-APS, (c) T4-APS, (d) T1-SPS, (e) T5-SHVOF, and (f) C-SPS, (g) TC-SPS, (h) A40%T-APS, (i) A13%T-APS, in which A denotes anatase phase, R denotes rutile phase, α is α-Al<sub>2</sub>O<sub>3</sub>, and γ is γ-Al<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 2
<p>Top surface 3D images of the coatings and R<sub>a</sub> provided by the confocal laser microscope.</p>
Full article ">Figure 3
<p>BSE FESEM micrographs of TiO<sub>2</sub> coatings’ cross-sections at two magnifications.</p>
Full article ">Figure 4
<p>Top surface SE FESEM micrographs of TiO<sub>2</sub> coatings.</p>
Full article ">Figure 5
<p>Top surface SE FESEM micrographs of T1-SPS and T5-SHVOF at high magnification, where (1) fully melted particles, (2) unmelted agglomerated particles, and (3) resolidified particles.</p>
Full article ">Figure 6
<p>FESEM micrographs of C-SPS (Cu<sub>2</sub>O) and TC-SPS (TiO<sub>2</sub>-Cu<sub>2</sub>O) coatings, left side: BSE cross-sectioned, right side: SE top surface images.</p>
Full article ">Figure 7
<p>FESEM micrographs of TiO<sub>2</sub>-Al<sub>2</sub>O<sub>3</sub> coatings, left side: BSE cross-sectioned, right side: SE top surface images.</p>
Full article ">Figure 8
<p>FESEM-EDX mapping of the cross-sectional view of TiO<sub>2</sub>-Cu<sub>2</sub>O and TiO<sub>2</sub>-Al<sub>2</sub>O<sub>3</sub> coatings.</p>
Full article ">Figure 9
<p>Antiviral activity assessment of TiO<sub>2</sub> coatings under (<b>a</b>) ambient light, (<b>b</b>) UVA light. Cu-10 min and SS-10 min were calculated by linear interpolation.</p>
Full article ">Figure 10
<p>The color difference between TiO<sub>2</sub> coatings produced by various thermal spray processes.</p>
Full article ">Figure 11
<p>Antiviral activity assessment of TiO<sub>2</sub>-Cu<sub>2</sub>O coatings under (<b>a</b>) ambient light, (<b>b</b>) UVA light, and (<b>c</b>) dark. Cu-10 min and SS-10 min were calculated by linear interpolation.</p>
Full article ">Figure 12
<p>Antiviral activity assessment of Al<sub>2</sub>O<sub>3</sub>-TiO<sub>2</sub> coatings under (<b>a</b>) ambient light, (<b>b</b>) UVA light. Cu-10 min and SS-10 min were calculated by linear interpolation.</p>
Full article ">Figure 13
<p>SEM micrographs of TiO<sub>2</sub> powder (Metco 102), Al<sub>2</sub>O<sub>3</sub>-40%TiO<sub>2</sub>, and Al<sub>2</sub>O<sub>3</sub>-40%TiO<sub>2</sub> are taken from the data sheets provided by Metco Oerlikon.</p>
Full article ">Figure 14
<p>Particle size distribution of the suspensions: (<b>a</b>) TiO<sub>2</sub>: d10 = 0.2 µm, d50 = 0.39 µm, d90 = 0.75 µm, (<b>b</b>) Cu<sub>2</sub>O: d10 = 2.9 µm, d50 = 5.9 µm, d90 = 11 µm, (<b>c</b>) TiO<sub>2</sub>-Cu<sub>2</sub>O: d10 = 0.2 µm, d50 = 0.68 µm, d90 = 2.8 µm, to deposit T1-SPS/T5-SHVOF, C-SPS, and TC-SPS coatings, respectively.</p>
Full article ">Figure 15
<p>Schematic of the antiviral activity assessment process.</p>
Full article ">
19 pages, 3397 KiB  
Article
Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture
by Muhammad Tahir, Azmat Ali Khan, Abdullah Bafaqeer, Naveen Kumar, Mohammad Siraj and Amanullah Fatehmulla
Catalysts 2023, 13(7), 1140; https://doi.org/10.3390/catal13071140 - 22 Jul 2023
Cited by 9 | Viewed by 1751
Abstract
Photocatalytic reduction of CO2 with CH4 through the dry reforming of methane (DRM) is an attractive approach to recycling greenhouse gases into valuable chemicals and fuels; however, this process is quite challenging. Although there is growing interest in designing efficient photocatalysts, [...] Read more.
Photocatalytic reduction of CO2 with CH4 through the dry reforming of methane (DRM) is an attractive approach to recycling greenhouse gases into valuable chemicals and fuels; however, this process is quite challenging. Although there is growing interest in designing efficient photocatalysts, they are less stable, and have lower photoactivity when employed for DRM reactions. Herein, we developed a noble metal-free hierarchical graphitic carbon nitride (HC3N4) loaded with cobalt (Co) for highly efficient and stable photocatalytic dry reforming of methane to produce synthesis gases (CO and H2). The performance of the newly designed Co/HC3N4 composite was tested for different reforming systems such as the dry reforming of methane, bi-reforming of methane (BRM) and reforming of CO2 with methanol–water. The performance of HC3N4 was much higher compared to bulk g-C3N4, whereas Co/HC3N4 was found to be promising for higher charge carrier separation and visible light absorption. The yield of CO and H2 with HC3N4 was 1.85- and 1.81-fold higher than when using g-C3N4 due to higher charge carrier separation. The optimized 2% Co/HC3N4 produces CO and H2 at an evolution rate of 555 and 41.2 µmol g−1 h−1, which was 18.28- and 1.74-fold more than using HC3N4 during photocatalytic dry reforming of methane (DRM), with a CH4/CO2 feed ratio of 1.0. This significantly enhanced photocatalytic CO and H2 evolution during DRM was due to efficient charge carrier separation in the presence of Co. The CH4/CO2 feed ratio was further investigated, and a 2:1 ratio was best for CO production. In contrast, the highest H2 was produced with a 1:1 feed ratio due to the competitive adsorption of the reactants over the catalyst surface. The performance of the composite was further investigated for bi-reforming methane and methanol. Using photocatalytic CO2 reduction with CH4/H2O, the production of CO and H2 was reduced, whereas significantly higher CO and H2 evolved using the BRM process involving methanol. Using methanol with CO2 and H2O, 10.77- and 1.39-fold more H2 and CO efficiency was achieved than when using dry reforming of methane. The composite was also very stable for continuous synthesis gas production during DRM in consecutive cycles. Thus, a co-assisted g-C3N4 nanotexture is promising for promoting photocatalytic activity and can be further explored in other solar energy applications. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) XRD analysis of g-C<sub>3</sub>N<sub>4</sub>, HC<sub>3</sub>N<sub>4</sub>, Co/HC<sub>3</sub>N<sub>4</sub>, (<b>b</b>) UV-vis DRS analysis of g-C<sub>3</sub>N<sub>4</sub>, HC<sub>3</sub>N<sub>4</sub>, Co/HC<sub>3</sub>N<sub>4</sub>, (<b>c</b>) PL analysis of g-C<sub>3</sub>N<sub>4</sub>, HC<sub>3</sub>N<sub>4</sub>, Co/HC<sub>3</sub>N<sub>4</sub>.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) g-C<sub>3</sub>N<sub>4</sub>, (<b>b</b>) HC<sub>3</sub>N<sub>4</sub>, (<b>c</b>) Co/HC<sub>3</sub>N<sub>4</sub>; (<b>d</b>) EDX mapping of Co/HC<sub>3</sub>N<sub>4</sub>, (<b>e</b>–<b>h</b>) color images showing the distribution of Co, C, N, and O over the Co/HC<sub>3</sub>N<sub>4</sub> surface.</p>
Full article ">Figure 3
<p>XPS analysis of Co/HC<sub>3</sub>N<sub>4</sub> for (<b>a</b>) N 1s, (<b>b</b>) C 1s, (<b>c</b>) Co 2p.</p>
Full article ">Figure 4
<p>Photocatalytic CO<sub>2</sub> reduction with CH<sub>4</sub> over g-C<sub>3</sub>N<sub>4</sub>, HC<sub>3</sub>N<sub>4</sub>, and Co-doped HC<sub>3</sub>N<sub>4</sub> composite samples with a CO<sub>2</sub>/CH<sub>4</sub> ratio of 1.0.</p>
Full article ">Figure 5
<p>Effect of CH<sub>4</sub>/CO<sub>2</sub> feed ratio in a photocatalytic DRM process over a Co/HC<sub>3</sub>N<sub>4</sub> composite on (<b>a</b>) CO evolution and (<b>b</b>) H<sub>2</sub> evolution.</p>
Full article ">Figure 5 Cont.
<p>Effect of CH<sub>4</sub>/CO<sub>2</sub> feed ratio in a photocatalytic DRM process over a Co/HC<sub>3</sub>N<sub>4</sub> composite on (<b>a</b>) CO evolution and (<b>b</b>) H<sub>2</sub> evolution.</p>
Full article ">Figure 6
<p>Photocatalytic CO<sub>2</sub> reduction through bi-reforming reactions: (<b>a</b>) CO<sub>2</sub> reduction with CH<sub>4</sub> and H<sub>2</sub>O and (<b>b</b>) CO<sub>2</sub> reduction with methanol and water.</p>
Full article ">Figure 7
<p>(<b>a</b>) Photocatalytic dry reforming of methane (DRM) over 2% Co/HC<sub>3</sub>N<sub>4</sub> composite per 4 h in consecutive three cycles; (<b>b</b>) XRD analysis of fresh and used 2% Co/HC<sub>3</sub>N<sub>4</sub> composite; (<b>c</b>) FTIR analysis of fresh and used 2% Co/HC<sub>3</sub>N<sub>4</sub> composite.</p>
Full article ">Figure 8
<p>Proposed mechanism for the photocatalytic dry and bi-reforming of methane to produce CO and H<sub>2</sub> over a Co/HC<sub>3</sub>N<sub>4</sub> composite.</p>
Full article ">
15 pages, 2665 KiB  
Article
Comparison of Photocatalytic Biocidal Activity of TiO2, ZnO and Au/ZnO on Escherichia coli and on Aspergillus niger under Light Intensity Close to Real-Life Conditions
by Mohamad Al Hallak, Thomas Verdier, Alexandra Bertron, Kevin Castelló Lux, Ons El Atti, Katia Fajerwerg, Pierre Fau, Julie Hot, Christine Roques and Jean-Denis Bailly
Catalysts 2023, 13(7), 1139; https://doi.org/10.3390/catal13071139 - 22 Jul 2023
Cited by 5 | Viewed by 2568
Abstract
Microbial contamination of the surface of building materials and subsequent release of microbial particles into the air can significantly affect indoor air quality. Avoiding the development or, at least, reducing the quantity of microorganisms growing on building materials is a key point to [...] Read more.
Microbial contamination of the surface of building materials and subsequent release of microbial particles into the air can significantly affect indoor air quality. Avoiding the development or, at least, reducing the quantity of microorganisms growing on building materials is a key point to reduce health risks for building occupiers. In that context, the antimicrobial activity of TiO2, ZnO and Au/ZnO was assessed by measuring log reductions of Escherichia coli and Aspergillus niger populations both in the dark and under a light intensity close to real-life conditions. The bactericidal activities (≥2.3 log reduction) of tested products were stronger than their fungicidal activities (≤1.4 log reduction) after 2 h of contact. Different parameters including concentration of photocatalyst, intensity of light (dark vs. 5 W/m2 UV-A), and duration of contact between photocatalyst and microbial cells and spores were investigated. Results of this study confirmed bactericidal activities of TiO2, ZnO and AuZnO on E. coli and brought new insight on their fungicidal activity on the spores of A. niger. They also confirmed the greatest antimicrobial efficiency of ZnO compared to TiO2 and its increased photocatalytic activity when decorated with Au, leading to the highest log reductions detected after 2 h of contact for both tested microorganisms (4 and 1.4 for E. coli and A. niger, respectively). The antimicrobial activity was enhanced by the duration of contact between microorganisms and nanoparticles of the different tested photocatalytic products. Full article
(This article belongs to the Special Issue Photocatalytic Nanomaterials for Abatement of Microorganisms)
Show Figures

Figure 1

Figure 1
<p>X-ray diffractograms of (<b>a</b>) TiO<sub>2</sub> P25 and (<b>b</b>) ZnO and Au/ZnO. Peaks were indexed by using the Inorganic Crystal Structure Database (ICSD).</p>
Full article ">Figure 2
<p>HRTEM images of Au/ZnO highlighting various morphologies of Au NPs (<b>a</b>,<b>b</b>); HRTEM images of Au/ZnO interface (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 3
<p>Average log reduction of <span class="html-italic">E. coli</span> 53126 after contact with different photocatalysts TiO<sub>2</sub> (blue bars), ZnO (orange bars) and AuZnO (grey bars) at 10 g/L concentration in the dark (<b>a</b>) or under an exposition to a light intensity of 5 W/m<sup>2</sup> (<b>b</b>). Results are expressed as mean ± SD (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 4
<p>Comparison of time-dependent bactericidal activities (<span class="html-italic">E. coli</span> 53126) of photocatalysts used at 1 and 10 g/L. Orange curves: control tubes; grey curves: 10 g/L of photocatalyst; blue curves: 1 g/L of photocatalyst. TiO<sub>2</sub> in the dark (<b>a</b>), TiO<sub>2</sub> under light (<b>b</b>), ZnO in the dark (<b>c</b>), ZnO under light (<b>d</b>), Au/ZnO in the dark (<b>e</b>) and Au/ZnO under light (<b>f</b>). The results presented are expressed as mean ± SD of two independent experiments.</p>
Full article ">Figure 5
<p>Time-dependent average log reduction of <span class="html-italic">A. niger</span> CBS 733.88 after contact TiO<sub>2</sub> (blue bars), ZnO (orange bars) and AuZnO (grey bars) at 10 g/L concentration in the dark (<b>a</b>) or under an exposition to a light intensity of 5 W/m<sup>2</sup> (<b>b</b>). Results are expressed as mean ± SD (<span class="html-italic">n</span> = 2).</p>
Full article ">Figure 6
<p>Comparison of time-dependent fungicidal activities (<span class="html-italic">A. niger</span> CBS 733.88) of photocatalysts used at 1 and 10 g/L. Orange curves: control tubes; grey curves: 10 g/L of photocatalyst; blue curves: 1 g/L of photocatalyst. ZnO in the dark (<b>a</b>), ZnO under light (<b>b</b>), Au/ZnO in the dark (<b>c</b>) and Au/ZnO under light (<b>d</b>). The results presented are the mean ± SD of two independent experiments.</p>
Full article ">Figure 7
<p>Effect of 10 g/L of TiO<sub>2</sub> on spores of <span class="html-italic">A. niger</span> throughout a 24 h incubation period under light. Orange curve: control tube; Blue curve: test tube. One representative experiment is shown.</p>
Full article ">
14 pages, 3832 KiB  
Article
Utilization of Loaded Cobalt onto MCM-48 Mesoporous Catalyst as a Heterogeneous Reaction in a Fixed Bed Membrane Reactor to Produce Isomerization Product from n-Heptane
by Nisreen S. Ali, Issam K. Salih, Hamed N. Harharah, Hasan Sh. Majdi, Hussein G. Salih, Khairi R. Kalash, Ali Al-Shathr, Farah T. Al-Sudani, Mahir A. Abdulrahman, Jamal M. Alrubaye, Talib M. Albayati, Noori M. Saady and Sohrab Zendehboudi
Catalysts 2023, 13(7), 1138; https://doi.org/10.3390/catal13071138 - 22 Jul 2023
Cited by 23 | Viewed by 1733
Abstract
The use of catalytic membranes as microstructured reactors without a separative function has proved effective. High catalytic activity is possible with minimal mass transport resistances if the reactant mixture is pushed to flow through the pores of a membrane that has been impregnated [...] Read more.
The use of catalytic membranes as microstructured reactors without a separative function has proved effective. High catalytic activity is possible with minimal mass transport resistances if the reactant mixture is pushed to flow through the pores of a membrane that has been impregnated with catalyst. In this study, n-heptane (C7H16) was hydrocracked and hydro-isomerized within a plug-flow zeolitic catalytic membrane-packed bed reactor. The metallic cobalt (Co) precursor at 3 wt.% was loaded onto support mesoporous materials MCM-48 to synthesize heterogeneous catalysis. The prepared MCM-48 was characterized by utilizing characterization techniques such as scanning electron microscopy (SEM), X-ray diffraction (XRD), energy dispersive X-ray analysis (EDAX), Fourier transform infrared (FTIR), nitrogen adsorption–desorption isotherms, and the Brunauer–Emmett–Teller (BET) surface area. The structural and textural characteristics of MCM-48 after encapsulation with Co were also investigated. The analyses were performed before and after metal loading. According to the results, the 3 wt.% Co/MCM-48 of metallic catalyst in a fixed bed membrane reactor (MR) appears to have an excellent catalytic activity of ~83% during converting C7H16 at 400 °C, whereas a maximum selectivity was approximately ~65% at 325 °C. According to our findings, the synthesized catalyst exhibits an acceptable selectivity to isomers with multiple branches, while making low aromatic components. In addition, a good catalytic stability was noticed for this catalyst over the reaction. Use of 3 wt.% Co/MCM-48 catalyst led to the highest isomerization selectivity as well as n-heptane conversion. Therefore, the heterogeneous catalysis MCM-48 is a promising option/ alternative for traditional hydrocracking and hydro-isomerization processes. Full article
(This article belongs to the Section Catalytic Reaction Engineering)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffraction pattern of MCM-48 and 3 wt.% Co/MCM-48.</p>
Full article ">Figure 2
<p>Nitrogen adsorption desorption isotherms for MCM-48 and 3% Co/MCM-48.</p>
Full article ">Figure 3
<p>BJH pore size distribution of MCM-48 and 3 wt.% Co/MCM-48.</p>
Full article ">Figure 4
<p>FT-IR spectra of MCM-48 and 3% Co/MCM-48.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM image of MCM-48, (<b>b</b>) EDX image for pure MCM/48, (<b>c</b>) SEM image of 3% Co/MCM-48, and (<b>d</b>) EDAX image for 3% Co/MCM-48.</p>
Full article ">Figure 6
<p>Conversion of <span class="html-italic">n</span>-heptane for the tested catalytic membrane.</p>
Full article ">Figure 7
<p>Isomerization product selectivity for the tested catalytic membrane.</p>
Full article ">Figure 8
<p>Cracking product selectivity for the tested catalytic membrane.</p>
Full article ">
27 pages, 44211 KiB  
Review
Core–Shell Catalysts for Conventional Oxidation of Alcohols: A Brief Review
by Luís M. M. Correia, Maxim L. Kuznetsov and Elisabete C. B. A. Alegria
Catalysts 2023, 13(7), 1137; https://doi.org/10.3390/catal13071137 - 21 Jul 2023
Cited by 4 | Viewed by 1624
Abstract
This review highlights recent research on the application of core–shell structured materials as catalysts in the oxidation of alcohols to value-added products, such as benzaldehyde, acetophenone, benzophenone, cinnamaldehyde, and vanillin, among others. While the application of various unconventional energy inputs (such as microwave [...] Read more.
This review highlights recent research on the application of core–shell structured materials as catalysts in the oxidation of alcohols to value-added products, such as benzaldehyde, acetophenone, benzophenone, cinnamaldehyde, and vanillin, among others. While the application of various unconventional energy inputs (such as microwave and ultrasound irradiation) was reported, this paper focuses on conventional heating. The oxidation of homocyclic aromatic, heterocyclic aromatic, aliphatic, and alicyclic alcohols catalyzed by core–shell composite catalysts is addressed. This work also highlights some unique advantages of core–shell nanomaterial catalysis, namely the flexibility of combining individual functions for specific purposes as well as the effect of various parameters on the catalytic performance of these materials. Full article
(This article belongs to the Special Issue Metallic Nanoparticles and Metal-Mediated Synthesis in Catalysis)
Show Figures

Figure 1

Figure 1
<p>Types of core–shell structures based on morphology: conventional single core–shell (<b>a</b>), multicore–shell (<b>b</b>), core–multishell (<b>c</b>), and multicore–multishell (<b>d</b>) structures.</p>
Full article ">Figure 2
<p>TEM images of SiO<sub>2</sub>-MAPTMS@PMMA core–shell particles with both core and shell spherical shapes (<b>a</b>) and of SiO<sub>2</sub>-MAPTMS@PS core–shell particles with different core and shell shapes, spherical and flower, respectively (<b>b</b>). Reproduced from [<a href="#B4-catalysts-13-01137" class="html-bibr">4</a>] with permission from the Walter de Gruyter GmbH.</p>
Full article ">Figure 3
<p>Composites of the core–shell (<b>a</b>), hollow core–shell (<b>b</b>) and yolk–shell (<b>c</b>) types. Adapted from [<a href="#B10-catalysts-13-01137" class="html-bibr">10</a>] with permission from the Royal Society of Chemistry.</p>
Full article ">Figure 4
<p>Preparation of the nanostructured Pd/Fe<sub>3</sub>O<sub>4</sub>@mCeO<sub>2</sub> core–shell (CTAB-cetyltrimethylammonium bromide; Brij—polyethylene glycol monododecyl ether). Adapted from [<a href="#B51-catalysts-13-01137" class="html-bibr">51</a>] with permission from Springer Nature.</p>
Full article ">Figure 5
<p>Preparation of Fe<sub>3</sub>O<sub>4</sub>@P4VP@FeCl<sub>3</sub> core–shell.</p>
Full article ">Figure 6
<p>Proposed mechanism for benzyl alcohol oxidation. Reproduced with permission from [<a href="#B51-catalysts-13-01137" class="html-bibr">51</a>].</p>
Full article ">Figure 7
<p>Proposed mechanism for the catalytic oxidation of benzyl alcohol. Reproduced from [<a href="#B52-catalysts-13-01137" class="html-bibr">52</a>] with permission from the Royal Society of Chemistry.</p>
Full article ">Figure 8
<p>Proposed mechanism for the oxidation of vanillic alcohol to vanillin by Co(II/III)@ZnO structured composites.</p>
Full article ">Figure 9
<p>Proposed mechanism for the oxidation of 2,5-hydroxymethylfurfural to 5-hydroxymethyl-furan-2-carboxylic acid catalyzed by a D-CeO<sub>2</sub>@N/C@TiO<sub>2</sub> core-shell.</p>
Full article ">Figure 10
<p>Possible mechanisms for the selective aerobic oxidation of crotyl oxidation to crotonaldehyde by Au@Pd core-shell: (<b>a</b>) mechanism involving β-H elimination and reaction of O<sub>2</sub> with Pd-H species and (<b>b</b>) redox mechanism. Reproduced with permission from [<a href="#B63-catalysts-13-01137" class="html-bibr">63</a>]. Copyright 2011 American Chemical Society.</p>
Full article ">Scheme 1
<p>Oxidation of benzyl alcohol to benzaldehyde and benzoic acid catalyzed by Au@Pd bimetallic core–shell structures supported on SiO<sub>2</sub> [<a href="#B32-catalysts-13-01137" class="html-bibr">32</a>].</p>
Full article ">Scheme 2
<p>Oxidation of 1-phenylethanol to acetophenone catalyzed by the Fe<sub>2</sub>O<sub>3</sub>@Ni<sub>3</sub>Al-LDH@Au<sub>25</sub>-0.053 nanostructure.</p>
Full article ">Scheme 3
<p>Selective oxidation of 2-phenylethanol to 2-phenylacetaldehyde catalyzed by Fe<sub>3</sub>O<sub>4</sub>@SiO<sub>2</sub>@VO(ephedrine)<sub>2</sub>.</p>
Full article ">Scheme 4
<p>Oxidation of benzhydrol to benzophenone.</p>
Full article ">Scheme 5
<p>Oxidation of 4-chlorobenzhydrol to 4-chlorobenzophenone catalyzed by RuO<sub>2</sub>@ZrO<sub>2</sub> core–shell.</p>
Full article ">Scheme 6
<p>Oxidation of cinnamyl alcohol to cinnamaldehyde catalyzed by the γ-Fe<sub>2</sub>O<sub>3</sub>@Ni<sub>3</sub>Al-LDH@Au<sub>25</sub> composite.</p>
Full article ">Scheme 7
<p>Oxidation of furfuryl alcohol to furfural and furoic acid catalyzed by the Fe<sub>3</sub>O<sub>4</sub>@SiO<sub>2</sub>@VO(ephedrine)<sub>2</sub> composite.</p>
Full article ">Scheme 8
<p>Oxidation of 2,5-hydroxymethylfurfural (HMF) to 2,5-diformylfuran (DFF), 5-hydroxymethyl-furan-2-carboxylic acid (HMFCA), 5-formyl-2-furoic acid (FFCA), and 2,5-furandicarboxylic acid (FDCA).</p>
Full article ">Scheme 9
<p>Oxidation of crotyl alcohol to crotonaldehyde and crotonic acid.</p>
Full article ">Scheme 10
<p>Oxidation of 1-octanol to octanal and 2-octanol to octanone.</p>
Full article ">Scheme 11
<p>Oxidation of geraniol to geranial.</p>
Full article ">Scheme 12
<p>Oxidation of cyclohexanol to cyclohexanone.</p>
Full article ">Scheme 13
<p>Oxidation of 3-methylheptanol to 3-methylheptanone.</p>
Full article ">
13 pages, 6669 KiB  
Article
Efficient and Stable Degradation of Triazophos Pesticide by TiO2/WO3 Nanocomposites with S-Scheme Heterojunctions and Oxygen Defects
by Wen Li, Chunxu Chen, Renqiang Yang, Shuangli Cheng, Xiaoyu Sang, Meiwen Zhang, Jinfeng Zhang, Zhenghua Wang and Zhen Li
Catalysts 2023, 13(7), 1136; https://doi.org/10.3390/catal13071136 - 21 Jul 2023
Cited by 2 | Viewed by 1421
Abstract
The prevalent utilization of organophosphorus pesticides presents a profound risk to the global environment, necessitating the immediate development of a secure and reliable methodology to mitigate this hazard. Photocatalytic technology, through the generation of robust oxidizing free radicals by suitable catalysts, offers a [...] Read more.
The prevalent utilization of organophosphorus pesticides presents a profound risk to the global environment, necessitating the immediate development of a secure and reliable methodology to mitigate this hazard. Photocatalytic technology, through the generation of robust oxidizing free radicals by suitable catalysts, offers a viable solution by effectively oxidizing organophosphorus pesticides, thus preserving environmental well-being. In this study, we successfully synthesized TiO2/WO3 (TO/WO) nanocomposites characterized by oxygen defects and S-scheme heterojunctions, demonstrating superior photocatalytic activity in the degradation of triazophos. Notably, the 60-TO/WO nanocomposite, wherein the proportion of WO comprises 60% of the total, exhibited optimal photocatalytic degradation activity, achieving a degradation rate of 78% within 120 min, and demonstrating exceptional stability, maintaining impressive degradation activity across four cycles. This performance was notably superior to that of standalone TO and WO. The presence of oxygen defects in WO was corroborated by electron paramagnetic resonance (EPR) spectroscopy. The mechanism at the heterojunction of the 60-TO/WO nanocomposite, identified as an S-scheme, was also confirmed by EPR and theoretical computations. Oxygen defects expedite charge transfer and effectively enhance the photocatalytic reaction, while the S-scheme effectively segregates photogenerated electrons and holes, thereby optimizing the photocatalytic oxidation of triazophos. This study introduces a novel nanocomposite material, characterized by oxygen defects and the S-scheme heterojunction, capable of effectively degrading triazophos and promoting environmental health. Full article
(This article belongs to the Special Issue Advances in Photocatalysis and Electrocatalysis Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) The XRD spectra of the samples. (<b>b</b>) FT-IR spectra of the samples.</p>
Full article ">Figure 2
<p>SEM image of (<b>a</b>) WO, (<b>b</b>) TO, and (<b>c</b>) 60-TO/WO nanocomposites. TEM image of (<b>d</b>) WO, (<b>e</b>) TO, and (<b>f</b>) 60-TO/WO nanocomposites (the inset in <a href="#catalysts-13-01136-f002" class="html-fig">Figure 2</a>f is the HRTEM image of 60-TO/WO nanocomposite). EDS spectrum of (<b>g</b>) 60-TO/WO nanocomposites. HAADF and elemental mapping images of (<b>h</b>) 60-TO/WO nanocomposites.</p>
Full article ">Figure 3
<p>Comparison of (<b>a</b>) survey scan, (<b>b</b>) W 4f; (<b>c</b>) O 1s, (<b>d</b>) Ti 2p XPS curves of WO, TO, and 60-TO/WO nanocomposites.</p>
Full article ">Figure 4
<p>(<b>a</b>) N<sub>2</sub> adsorption-desorption isotherms and pore size distribution curves (insert) of TO, WO, and TO/WO nanocomposites; (<b>b</b>) The <span class="html-italic">S</span><sub>BET</sub> for the TO, WO, and TO/WO nanocomposites.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV-vis DRS of WO, TO, and TO/WO nanocomposites. (<b>b</b>) Tauc plots of (<span class="html-italic">αhν</span>)<sup>2</sup> versus energy (<span class="html-italic">hν</span>) of WO and TO.</p>
Full article ">Figure 6
<p>(<b>a</b>) The photocatalytic degradation of triazophos. (<b>b</b>) The histogram of photocatalytic degradation of triazophos. (<b>c</b>) The linear transform ln(C<sub>0</sub>/C) vs. times curves of triazophos degradation. (<b>d</b>) The photodegradation performance of photocatalysts over four cycles.</p>
Full article ">Figure 7
<p>(<b>a</b>) Photocurrent density–time curves of TO, WO, and 60-TO/WO nanocomposites. (<b>b</b>) Nyquist plots of TO, WO, and 60-TO/WO nanocomposites.</p>
Full article ">Figure 8
<p>(<b>a</b>) The schematic diagrams of charge transfer in Type-II and S-scheme heterojunction. (<b>b</b>) DMPO spin trapping EPR spectra of 60-TO/WO nanocomposites (·O<sub>2</sub><sup>−</sup> and ·OH). (<b>c</b>) Capture the EPR spectra of WO atomic defects. The work function of (<b>d</b>) WO (200) and (<b>e</b>) TO (101) facet. (<b>f</b>) Diagram of the formation process of the S-scheme at the TO/WO nanocomposites heterojunction.</p>
Full article ">Figure 9
<p>The photocatalysis mechanism of TO/WO nanocomposites under sunlight irradiation without co-catalysts.</p>
Full article ">Scheme 1
<p>Synthetic flowchart of TO/WO nanocomposites.</p>
Full article ">
23 pages, 7690 KiB  
Article
The Impact of Different Green Synthetic Routes on the Photocatalytic Potential of FeSnO2 for the Removal of Methylene Blue and Crystal Violet Dyes under Natural Sunlight Exposure
by Arifa Shaukat, Muhammad Akhyar Farrukh, Kok-Keong Chong, Rabia Nawaz, Muhammad Tariq Qamar, Shahid Iqbal, Nasser S. Awwad and Hala A. Ibrahium
Catalysts 2023, 13(7), 1135; https://doi.org/10.3390/catal13071135 - 21 Jul 2023
Cited by 6 | Viewed by 1870
Abstract
FeSnO2 nanocomposites were synthesized via the green method using aqueous leaf extracts of Lawsonia inermis and Phyllanthus embilica plants. The role of polyphenols based on reduction potentials for the synthesis of FeSnO2 was also highlighted. The synthesized materials were examined by [...] Read more.
FeSnO2 nanocomposites were synthesized via the green method using aqueous leaf extracts of Lawsonia inermis and Phyllanthus embilica plants. The role of polyphenols based on reduction potentials for the synthesis of FeSnO2 was also highlighted. The synthesized materials were examined by using TGA and DSC, FT-IR, XRD, and SEM with EDX analysis. Tetragonal rutile and distorted hexagonal structures were observed in SEM images of the FeSnO2 nanocomposites and compared with an FeSnO2 nanocomposite prepared using the sol-gel method. Scherer’s formula yielded crystallite sizes of 29.49, 14.54, and 20.43 nm; however, the average crystallite size assessed employing the Williamson–Hall equation was found to be 20.85, 11.30, and 14.86 nm by using the sol-gel and green techniques, using extracts from Lawsonia inermis and Phyllanthus embilica. The band gap was determined by using the Tauc and Wood equations, and photocatalytic activity was analyzed to determine the degradation of methylene blue (MB) and crystal violet (CV) under the illumination of natural sunlight. It was observed that the sample prepared by means of the green method using the leaf extract of Lawsonia inermis showed the best photocatalytic activity of 84%, with a particle size of 14.54 nm, a 3.10 eV band gap, and a specific surface area of 55.68 m2g−1. Full article
(This article belongs to the Special Issue Advanced Materials for Application in Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The comparison of FT-IR spectra of FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 2
<p>FTIR spectra of aqueous leaf extract of (<b>a</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>b</b>) <span class="html-italic">Phyllanthus embilica</span>.</p>
Full article ">Figure 3
<p>TGA curves of FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 4
<p>ln(1 − x) versus time plot for FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 5
<p>Redfern and Coats plot for FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 6
<p>XRD patterns of FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 7
<p>Bandgap evaluation of FeSnO<sub>2</sub> synthesized using the (<b>a</b>) sol-gel method and green met-ods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts using the Wood and Tauc plot.</p>
Full article ">Figure 8
<p>SEM images and EDX spectra of FeSnO<sub>2</sub> nanocomposite synthesized via the (<b>a</b>,<b>d</b>) sol-gel method and the (<b>b</b>,<b>e</b>) green method using <span class="html-italic">Lawsonia inermis</span> leaf extract and (<b>c</b>,<b>f</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extract.</p>
Full article ">Figure 9
<p>Percentage degradation of MB by FeSnO<sub>2</sub> photocatalysts synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts under exposure to natural sunlight. (<b>d</b>) Photolysis of MB.</p>
Full article ">Figure 10
<p>Percentage degradation of CV by FeSnO<sub>2</sub> photocatalysts synthesized using the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts under exposure to natural sunlight. (<b>d</b>) Photolysis of the CV.</p>
Full article ">Figure 11
<p>The comparison of the rate of removal of crystal violet (CV) dye by FeSnO<sub>2</sub> synthesized via the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 12
<p>The comparison of the rate of removal of methylene blue (MB) dye by FeSnO<sub>2</sub> synthesized via the (<b>a</b>) sol-gel method and green methods using (<b>b</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>c</b>) <span class="html-italic">Phyllanthus embillica</span> leaf extracts.</p>
Full article ">Figure 13
<p>FTIR spectra of (<b>a</b>) methylene blue and (<b>b</b>) crystal violet dyes after 60 and 120 min during the degradation process by FeSnO<sub>2</sub> synthesized using <span class="html-italic">Lawsonia inermis</span> leaf extract.</p>
Full article ">Figure 14
<p>Synthesis of polyphenol ions and the possible mechanism of the green synthesis of FeSnO<sub>2</sub> nanocomposites.</p>
Full article ">Figure 15
<p>High-resolution XPS spectra of (<b>a</b>) Fe 2p, (<b>b</b>) Sn 3d and (<b>c</b>) O 1s in the FeSnO<sub>2</sub> composite.</p>
Full article ">Figure 16
<p>The representative leaves of (<b>a</b>) <span class="html-italic">Lawsonia inermis</span> and (<b>b</b>) <span class="html-italic">Phyllanthus embilica</span>.</p>
Full article ">
15 pages, 5679 KiB  
Article
Photocatalytic Degradation of Ciprofloxacin with Supramolecular Materials Consisting of Nitrogenous Organic Cations and Metal Salts
by Chenfei Ren, Jian Li, Xingxing Zhang and Yunyin Niu
Catalysts 2023, 13(7), 1134; https://doi.org/10.3390/catal13071134 - 21 Jul 2023
Cited by 2 | Viewed by 1329
Abstract
The design and synthesis of composite materials with new structures/properties have important practical significance for the degradation of organic pollutants in aquatic environments. On this basis, five new supramolecular materials {[L1]2·[Cu4I8]}(1), {[L1 [...] Read more.
The design and synthesis of composite materials with new structures/properties have important practical significance for the degradation of organic pollutants in aquatic environments. On this basis, five new supramolecular materials {[L1]2·[Cu4I8]}(1), {[L1]2·[Ag4I8]}(2), {[L2]·[ZnBr4]}(3), {[L3]2·[AgI5]}(4), {[L3]·[CdBr3Cl]}(5) were synthesized by introducing an amino group into a series of nitrogen-containing cationic ligands (L1L3) through the reaction of polybromomethylbenzene with 4-aminopyridine. The degradation effect of catalysts 15 on ciprofloxacin (CIP) under visible light was studied using their potential catalytic properties. The results showed that compounds 1 and 4 had better degradation effects compared to other compounds. Moreover, compounds 1 and 4 were proved to be excellent catalysts for the photocatalytic degradation of CIP with cyclic experiments. Through further exploration, it was found that neutral conditions and 20 mg compound dosage were more conducive to the photodegradation of CIP by the compound. Through free radical capture experiments, it was found that ·OH played a major role in the photodegradation of CIP. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Structural formula of several common quinolone antibiotics.</p>
Full article ">Figure 2
<p>Reaction equation of ligands <b>L1</b>–<b>L3</b>.</p>
Full article ">Figure 3
<p>(<b>a</b>) Structural monomer diagram of compound <b>1</b>; (<b>b</b>) Stacking diagram of compound <b>1</b> in the c-direction.</p>
Full article ">Figure 4
<p>(<b>a</b>) Structural monomer diagram of compound <b>2</b>; (<b>b</b>) Stacking diagram of compound <b>2</b> in the c-direction.</p>
Full article ">Figure 5
<p>(<b>a</b>) Structural monomer diagram of compound <b>3</b>; (<b>b</b>) Stacking diagram of compound <b>3</b> in the b-direction.</p>
Full article ">Figure 6
<p>(<b>a</b>) Asymmetric structural units of compound <b>4</b>; (<b>b</b>) Stacking diagram of compound <b>4</b> in the a-direction.</p>
Full article ">Figure 7
<p>(<b>a</b>) Asymmetric structural units of compound <b>5</b>; (<b>b</b>) Stacking diagram of compound <b>5</b> in the a-direction.</p>
Full article ">Figure 8
<p>Band gap energy diagram of compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 9
<p>Rate chart of photocatalytic degradation of CIP by compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 10
<p>(<b>a</b>) Degradation effect of compound <b>1</b> on CIP at different dosages; (<b>b</b>) Degradation effect chart of compound <b>1</b> on CIP under different pH values.</p>
Full article ">Figure 11
<p>Effects of different free radical scavengers on the removal of CIP from compounds <b>1</b> (<b>a</b>) and <b>4</b> (<b>b</b>).</p>
Full article ">Figure 12
<p>Removal rate of CIP by compounds <b>1</b> (<b>a</b>) and <b>4</b> (<b>b</b>) in three cycle experiments.</p>
Full article ">Figure 13
<p>Infrared diagram of compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 14
<p>UV-VIS diffuse reflectance spectra of compounds <b>1</b>–<b>5</b>.</p>
Full article ">Figure 15
<p>PXRD diagram of compounds <b>1</b>–<b>5</b>.</p>
Full article ">
13 pages, 1933 KiB  
Article
A New Pseudomonas aeruginosa Isolate Enhances Its Unusual 1,3-Propanediol Generation from Glycerol in Bioelectrochemical System
by Julia Pereira Narcizo, Lucca Bonjy Kikuti Mancilio, Matheus Pedrino, María-Eugenia Guazzaroni, Adalgisa Rodrigues de Andrade and Valeria Reginatto
Catalysts 2023, 13(7), 1133; https://doi.org/10.3390/catal13071133 - 20 Jul 2023
Viewed by 1577
Abstract
The ability of some bacteria to perform Extracellular Electron Transfer (EET) has been explored in bioelectrochemical systems (BES) to obtain energy or chemicals from pure substances or residual substrates. Here, a new pyoverdine-producing Pseudomonas aeruginosa strain was isolated from an MFC biofilm oxidizing [...] Read more.
The ability of some bacteria to perform Extracellular Electron Transfer (EET) has been explored in bioelectrochemical systems (BES) to obtain energy or chemicals from pure substances or residual substrates. Here, a new pyoverdine-producing Pseudomonas aeruginosa strain was isolated from an MFC biofilm oxidizing glycerol, a by-product of biodiesel production. Strain EL14 was investigated to assess its electrogenic ability and products. In an open circuit system (fermentation system), EL14 was able to consume glycerol and produce 1,3-propanediol, an unusual product from glycerol oxidation in P. aeruginosa. The microbial fuel cell (MFC) EL14 reached a current density of 82.4 mA m−2 during the first feeding cycle, then dropped sharply as the biofilm fell off. Cyclic voltammetry suggests that electron transfer to the anode occurs indirectly, i.e., through a redox substance, with redox peak at 0.22 V (vs Ag/AgCl), and directly probably by membrane redox proteins, with redox peak at 0.05 V (vs Ag/AgCl). EL14 produced added-value bioproducts, acetic and butyric acids, as well as 1,3 propanediol, in both fermentative and anodic conditions. However, the yield of 1,3-PDO from glycerol was enhanced from 0.57 to 0.89 (mol of 1,3-PDO mol−1 of glycerol) under MFC conditions compared to fermentation. This result was unexpected, since successful 1,3-PDO production is not usually associated with P. aeruginosa glycerol metabolism. By comparing EL14 genomic sequences related to the 1,3-PDO biosynthesis with P. aeruginosa reference strains, we observed that strain EL14 has three copies of the dhaT gene (1,3-propanediol dehydrogenase a different arrangement compared to other Pseudomonas isolates). Thus, this work functionally characterizes a bacterium never before associated with 1,3-PDO biosynthesis, indicating its potential for converting a by-product of the biodiesel industry into an emerging chemical product. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Current density (black line) produced in the MFC with an external resistor of 2200 Ω and glycerol concentration of 22 mM. The shadows represent the standard deviation in the replicates.</p>
Full article ">Figure 2
<p>Cyclic voltammograms of the anode with planktonic cells (red line) and of the anode with biofilm in fresh and sterile medium (blue line). Abiotic control is represented by the black line. ν = 2 mV s<sup>−1</sup>, pH 6.9.</p>
Full article ">Figure 3
<p>Concentration of glycerol (gray line), 1,3-PDO (red line), acetic acid (blue line), and butyric acid (green line), and the current density (black line) produced by the MFC with an external 2200 Ω resistor (<b>A</b>) and in the fermentative system over time (<b>B</b>). Y<sub>P/S</sub> conversion factor from glycerol to 1,3-PDO (orange line), acetic acid (purple line), and butyric acid (pink line) during MFC operation (<b>C</b>) and in the fermentation (<b>D</b>). Optical density at 600 nm (gray line) and pH (red line) in MFC (<b>E</b>) and fermentation (<b>F</b>).</p>
Full article ">Figure 4
<p>Metabolism glycerol in natural 1,3-PDO producers. Pyruvate is converted to different organic compounds depending on the microorganism. 3-HPA, 3-hydroxypropionaldehyde; DHA, dihydroxyacetone; DHAP, dihydroxyacetone phosphate; PEP, phosphoenolpyruvate.</p>
Full article ">Figure 5
<p>Protein sequence alignment of the three 1,3-propanediol dehydrogenases variants (DhaT) identified in the EL14 strain genome and compared with DhaT from <span class="html-italic">P. aeruginosa</span> PAO1 (brownish bar). All mismatches among the protein sequences are represented by blank spaces. The upper black bar represents the number of amino acids in the protein sequences.</p>
Full article ">
15 pages, 5212 KiB  
Article
Templating Synthesis of Hierarchically Porous Carbon with Magnesium Salts for Electrocatalytic Reduction of 4-Nitrophenol
by Wanyi Gan, Ping Xiao and Junjiang Zhu
Catalysts 2023, 13(7), 1132; https://doi.org/10.3390/catal13071132 - 20 Jul 2023
Cited by 1 | Viewed by 1431
Abstract
Hierarchically porous carbon (PC) was synthesized by a templating method, using magnesium salts (Mg(HCO3)2, MgC2O4 and MgO) as template precursors and citric acid as carbon precursor. During the carbonization process, besides the production of MgO particles, [...] Read more.
Hierarchically porous carbon (PC) was synthesized by a templating method, using magnesium salts (Mg(HCO3)2, MgC2O4 and MgO) as template precursors and citric acid as carbon precursor. During the carbonization process, besides the production of MgO particles, many gases (e.g., CO2/NO2/H2O) were also released and acted as a porogen to generate pores in carbon. The resulting composite (MgO@C) was subsequently treated with HCl solution to remove the MgO templates, yielding hierarchically porous carbon. The surface oxygen functional groups over porous carbon were characterized by TPD and XPS, which showed that the PC-bic, synthesized using Mg(HCO3)2 as the template precursor, had the highest value among the PCs. As expected, the PC-bic exhibited the best performances for electrocatalytic reduction of 4-nitrophenol, with a peak current of −135.5 μA at −0.679 V. The effects of 4-nitrophenol concentration, buffer solution pH and scanning rate on the electrocatalytic activities, as well as the stability of PC-bic for the reaction were investigated. Full article
(This article belongs to the Special Issue Graphene Related Materials for Catalytic Applications)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>A</b>) the MgO@C-<span class="html-italic">n</span> composites, and (<b>B</b>) the PCs.</p>
Full article ">Figure 2
<p>TGA and DTG profiles of (<b>A</b>) citric acid, (<b>B</b>) Mg(HCO<sub>3</sub>)<sub>2</sub> + citric acid, (<b>C</b>) MgC<sub>2</sub>O<sub>4</sub> + citric acid and (<b>D</b>) MgO + citric acid conducted in N<sub>2</sub> atmosphere.</p>
Full article ">Figure 3
<p>TGA profiles of (<b>A</b>) the MgO@C composites and (<b>B</b>) the PCs conducted in air atmosphere.</p>
Full article ">Figure 4
<p>TEM images of (<b>A</b>,<b>B</b>) MgO@C-bic and (<b>D</b>,<b>E</b>) PC-bic, and STEM elemental mapping of the Mg, O and C atoms of (<b>C</b>) MgO@C-bic and (<b>F</b>) PC-bic.</p>
Full article ">Figure 5
<p>N<sub>2</sub> physisorption isotherms and pore distributions of (<b>A</b>,<b>B</b>) the MgO@C composites and (<b>C</b>,<b>D</b>) the PCs.</p>
Full article ">Figure 6
<p>MS signal of (<b>A</b>) CO and (<b>B</b>) CO<sub>2</sub> obtained from TPD measurements over the PC-<span class="html-italic">n</span> (<span class="html-italic">n</span> = bic, oxa and oxi).</p>
Full article ">Figure 7
<p>Different types of oxygen based on the deconvoluted O 1s XPS spectra of PC-bic and PC-oxa.</p>
Full article ">Figure 8
<p>CV curves of the PCs for electrocatalytic reduction of 4-nitrophenol (0.1 mM) in 0.2 M phosphate buffer solution at pH 6.0, with a scan rate of 20 mV s<sup>−1</sup>.</p>
Full article ">Figure 9
<p>(<b>A</b>) Plot of peak current vs. concentrations of 4-nitrophenol; (<b>B</b>) Plot of reduction peak current and potential vs. pH value; (<b>C</b>) Plot of reduction peak current vs. square root of scan rate.</p>
Full article ">Scheme 1
<p>The electrochemical reduction mechanism of 4-nitrophenol.</p>
Full article ">
31 pages, 3845 KiB  
Article
Educational Scale-Bridging Approach towards Modelling of Electric Potential, Electrochemical Reactions, and Species Transport in PEM Fuel Cell
by Ambrož Kregar, Klemen Zelič, Andraž Kravos and Tomaž Katrašnik
Catalysts 2023, 13(7), 1131; https://doi.org/10.3390/catal13071131 - 20 Jul 2023
Cited by 1 | Viewed by 1540
Abstract
The use of hydrogen fuel cells as a mobile source of electricity could prove key to the future decarbonisation of heavy-duty road and marine transportation. Due to the complex interplay of various physicochemical processes in fuel cells, further development of these devices will [...] Read more.
The use of hydrogen fuel cells as a mobile source of electricity could prove key to the future decarbonisation of heavy-duty road and marine transportation. Due to the complex interplay of various physicochemical processes in fuel cells, further development of these devices will depend on concerted efforts by researchers from various fields, who often lack in-depth knowledge of different aspects of fuel cell operation. These knowledge gaps can be filled by information that is scattered in a wide range of literature, but is rarely covered in a concise and condensed manner. To address this issue, we propose an educational-scale-bridging approach towards the modelling of most relevant processes in the fuel cell that aims to adequately describe the causal relations between the processes involved in fuel cell operation. The derivation of the model equations provides an intuitive understanding of the electric and chemical potentials acting on protons at the microscopic level and relates this knowledge to the terminology commonly used in fuel cell research, such as catalyst electric overpotential and internal membrane resistance. The results of the model agreed well with the experimental data, indicating that the proposed simple mathematical description is sufficient for an intuitive understanding of fuel cell operation. Full article
(This article belongs to the Topic Hydrogen Energy Technologies, 2nd Volume)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic overview of the fuel cell cross-section with the description model variables (red) and boundary conditions (blue). The dimensions of the individual fuel cell components are not shown to scale.</p>
Full article ">Figure 2
<p>Schematic representation of the configuration of the particles, involved in hydrogen reduction and oxidation. In initial State 1, both protons (blue circles) reside in the bulk electrolyte, screened from the electrode by the electric double layer (EDL). Proton reduction occurs as the protons are first adsorbed on the electrode surface (States 2 and 3). With sufficient energy, the protons can form a transition state (4), which chemically binds into a hydrogen molecule (State 5) and finally detaches from the electrode surface (State 6). When the hydrogen molecule is oxidised, the reaction proceeds in the opposite direction from State 6 to State 1.</p>
Full article ">Figure 3
<p>Schematic representation of the energy levels of the states, involved in the hydrogen electrochemical reaction on a electrically neutral (<b>a</b>) and a charged (<b>b</b>) electrode. The numbering of the states relates to <a href="#catalysts-13-01131-f002" class="html-fig">Figure 2</a>: 1: <math display="inline"><semantics><mrow><mn>2</mn><msup><mi>H</mi><mo>+</mo></msup><mo>+</mo><mn>2</mn><msup><mi>e</mi><mo>−</mo></msup></mrow></semantics></math>, 2: <math display="inline"><semantics><mrow><msup><mi>H</mi><mo>+</mo></msup><mo>+</mo><msubsup><mi>H</mi><mrow><mi>a</mi><mi>d</mi><mi>s</mi></mrow><mo>+</mo></msubsup><mo>+</mo><mn>2</mn><msup><mi>e</mi><mo>−</mo></msup></mrow></semantics></math>, 3: <math display="inline"><semantics><mrow><mn>2</mn><msubsup><mi>H</mi><mrow><mi>a</mi><mi>d</mi><mi>s</mi></mrow><mo>+</mo></msubsup><mo>+</mo><mn>2</mn><msup><mi>e</mi><mo>−</mo></msup></mrow></semantics></math>, 4: <math display="inline"><semantics><mrow><msubsup><mi>H</mi><mrow><mn>2</mn><mo>,</mo><mi>a</mi><mi>d</mi><mi>s</mi></mrow><mrow><mo>+</mo><mn>2</mn></mrow></msubsup><mo>+</mo><mn>2</mn><msup><mi>e</mi><mo>−</mo></msup></mrow></semantics></math>, 5: <math display="inline"><semantics><msub><mi>H</mi><mrow><mn>2</mn><mo>,</mo><mi>a</mi><mi>d</mi><mi>s</mi></mrow></msub></semantics></math>, 6: <math display="inline"><semantics><msub><mi>H</mi><mn>2</mn></msub></semantics></math>. On the neutral electrode (<b>a</b>), the electrochemical enthalpy levels (purple) are defined only by the chemical enthalpy (orange) with no electric contribution (green). In this case, the adsorption of protons to the electrode surface (<math display="inline"><semantics><mrow><mn>1</mn><mo>→</mo><mn>2</mn><mo>→</mo><mn>3</mn></mrow></semantics></math>) does not affect the enthalpy, which is increased only when the transition state (4) is formed. The formation of a hydrogen molecule (<math display="inline"><semantics><mrow><mn>4</mn><mo>→</mo><mn>5</mn></mrow></semantics></math>) results in a significant decrease in the chemical enthalpy. When the electrode is charged (<b>b</b>), proton adsorption requires their transition through the EDL, which increases the electric and electrochemical energy of the system. As a result, the electrochemical enthalpy of the initial state 0 decreases by <math display="inline"><semantics><mrow><mo>−</mo><mn>2</mn><msub><mi>e</mi><mn>0</mn></msub><mi>U</mi></mrow></semantics></math> and of the transition State 4 by <math display="inline"><semantics><mrow><mo>−</mo><mn>2</mn><mi>α</mi><msub><mi>e</mi><mn>0</mn></msub><mi>U</mi></mrow></semantics></math>, where <math display="inline"><semantics><mi>α</mi></semantics></math> is the charge transfer coefficient. The change in enthalpy levels affects the share of particles that have sufficient thermal energy to overcome the enthalpy barrier to the transition state (red dashed area) and, thus, affects the rates of the electrochemical reaction.</p>
Full article ">Figure 4
<p>Electrochemical enthalpy landscape of the reaction Equation (<a href="#FD2-catalysts-13-01131" class="html-disp-formula">2</a>) with electric potential close to equilibrium. The electric charging results in the electrochemical enthalpy of the initial and the final states to be similar, which in turn leads to similar transition state barriers in both directions. Small deviations between enthalpy levels can be due to the different activities of the reactants and products, which can promote the reaction rate toward reduction or oxidation.</p>
Full article ">Figure 5
<p>Schematic representation of the catalyst layer. Catalyst nanoparticles (white) are dispersed on the surface of the carbon support structure (black) with the volumetric ratio <math display="inline"><semantics><msub><mi>μ</mi><mi>C</mi></msub></semantics></math>. The catalyst and carbon support are covered with ionomer film (light blue) with the volumetric ratio <math display="inline"><semantics><msub><mi>μ</mi><mrow><mi>i</mi><mi>o</mi><mi>n</mi></mrow></msub></semantics></math>, which allows the protons to travel to the catalyst surface. The gaseous species (hydrogen, oxygen) reside in the void volume of the catalyst (white, volumetric ration <math display="inline"><semantics><msub><mi>μ</mi><mn>0</mn></msub></semantics></math>).</p>
Full article ">Figure 6
<p>Schematic representation of the enthalpy levels of the states involved in the electrochemical reaction in the fuel cell cathode. Reduction (solid line) and oxidation (dashed line) can proceed between the initial state (<span class="html-italic">init</span>) and the final state (<span class="html-italic">fin</span>) via different transition states (<span class="html-italic">trans,Red</span> and <span class="html-italic">trans,Ox</span>) with different chemical enthalpies (orange), which are affected differently by electric charging (green), resulting in different electrochemical enthalpy barriers (purple) for the reduction and oxidation reactions.</p>
Full article ">Figure 7
<p>Schematic presentation of the electric field and potential in the fuel cell ionomer. The imbalance between the concentrations of protons (red) and sulphonic groups (dashed orange) in the anode and cathode layers results in a net positive charge density at the anode and a negative charge density at the cathode (blue). This creates an electric field within the membrane (purple) that forces the protons to move through the membrane and also leads to a difference in the ionomer electric potential (green) between the anode and cathode.</p>
Full article ">Figure 8
<p>Comparison between measured polarisation curves and data produced by the model.</p>
Full article ">Figure 9
<p>Comparison between the different contributions to the fuel cell voltage drop in Equation (<a href="#FD79-catalysts-13-01131" class="html-disp-formula">79</a>) at different gas pressures: 1 bar (black), 1.4 bar (blue), 2.0 bar (green), and 2.5 bar (red). The contributions are: anode overpotential <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>a</mi><mi>n</mi></mrow></msub></semantics></math> (dashed), cathode overpotential <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>c</mi><mi>a</mi><mi>t</mi></mrow></msub></semantics></math> (solid), and membrane voltage drop <math display="inline"><semantics><msub><mi>U</mi><mrow><mi>m</mi><mi>e</mi><mi>m</mi></mrow></msub></semantics></math> (dotted). Note that the anode and cathode overpotentials depend strongly on the gas pressure, which changes the concentration of the reactants. Since the membrane conductivity is assumed to be constant, the membrane voltage drop is the same at all pressures.</p>
Full article ">Figure 10
<p>Concentrations of reactants and products in the fuel cell catalyst layers at different current densities and gas inlet pressures: hydrogen in anode catalyst (dashed), oxygen in cathode catalyst (solid), and water in cathode catalyst (dotted linen). Increasing the current density results in a lower concentration of reactants (hydrogen and oxygen) due to electrochemical consumption and a higher concentration of product (water).</p>
Full article ">Figure 11
<p>Schematic presentation of the proton enthalpy in the fuel cell electrode–membrane assembly during operation at two different current densities: (<b>a</b>) <math display="inline"><semantics><mrow><msub><mi>j</mi><mrow><mi>e</mi><mi>l</mi></mrow></msub><mo>=</mo><mn>0</mn></mrow></semantics></math> and (<b>b</b>) <math display="inline"><semantics><mrow><msub><mi>j</mi><mrow><mi>e</mi><mi>l</mi></mrow></msub><mo>=</mo><msub><mi>j</mi><mrow><mi>r</mi><mi>e</mi><mi>f</mi></mrow></msub></mrow></semantics></math>. The total electrochemical enthalpy (purple) can be divided into chemical (orange) and electrical (green) contributions. Protons in the “Anode” are bound into a hydrogen molecule with low chemical enthalpy (<math display="inline"><semantics><mrow><mo>Δ</mo><msub><mi>H</mi><mrow><mi>c</mi><mi>h</mi><mo>,</mo><mi>H</mi><mi>O</mi><mi>R</mi></mrow></msub></mrow></semantics></math>) and large electric energy (<math display="inline"><semantics><msub><mi>U</mi><mrow><mi>a</mi><mi>n</mi></mrow></msub></semantics></math>). The transition to the “Membrane” requires the splitting of the molecule into two protons and two electrons, which increases the chemical enthalpy, but reduces the electric energy due to electrode charging. The reaction with oxygen at the “Cathode” results in the binding of protons into a water molecule with a large decrease in chemical energy (<math display="inline"><semantics><mrow><mo>Δ</mo><msub><mi>H</mi><mrow><mi>c</mi><mi>h</mi><mo>,</mo><mi>O</mi><mi>R</mi><mi>R</mi></mrow></msub></mrow></semantics></math>) and an accompanying increase in electric energy (<math display="inline"><semantics><msub><mi>U</mi><mrow><mi>c</mi><mi>a</mi><mi>t</mi></mrow></msub></semantics></math>), which causes a net electric potential difference between the cathode and anode (<math display="inline"><semantics><msub><mi>U</mi><mrow><mi>F</mi><mi>C</mi></mrow></msub></semantics></math>). Under OCV conditions without electric current (<b>a</b>), the electrode overpotentials <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>a</mi><mi>n</mi></mrow></msub></semantics></math> and <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>c</mi><mi>a</mi><mi>t</mi></mrow></msub></semantics></math> are determined by the Nernst Equation (<a href="#FD25-catalysts-13-01131" class="html-disp-formula">25</a>), while the potential drop across the membrane is zero. When current flows through the cell (<b>b</b>), larger overpotentials <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>a</mi><mi>n</mi></mrow></msub></semantics></math> and <math display="inline"><semantics><msub><mi>η</mi><mrow><mi>c</mi><mi>a</mi><mi>t</mi></mrow></msub></semantics></math> are needed to provide a sufficient rate of electrochemical reactions. Transport through the membrane requires an electric potential difference <math display="inline"><semantics><msub><mi>U</mi><mrow><mi>m</mi><mi>e</mi><mi>m</mi></mrow></msub></semantics></math> across the membrane.</p>
Full article ">Figure A1
<p>Schematic representation of the energy distribution function <math display="inline"><semantics><mrow><mi>w</mi><mo>(</mo><mi>E</mi><mo>)</mo></mrow></semantics></math> and the number of particles with sufficient energy for the reaction.</p>
Full article ">Figure A2
<p>Schematic representation of the energy distribution function <math display="inline"><semantics><mrow><mi>w</mi><mo>(</mo><mi>E</mi><mo>)</mo></mrow></semantics></math> (red line) and the probability that a particle has enough energy for a reaction (dashed red area).</p>
Full article ">
18 pages, 4812 KiB  
Article
Polymer-Grafted 3D-Printed Material for Enzyme Immobilization—Designing a Smart Enzyme Carrier
by Daniela Eixenberger, Aditya Kumar, Saskia Klinger, Nico Scharnagl, Ayad W. H. Dawood and Andreas Liese
Catalysts 2023, 13(7), 1130; https://doi.org/10.3390/catal13071130 - 20 Jul 2023
Cited by 4 | Viewed by 1504
Abstract
One way to enhance the flow properties of packed bed reactors, including efficient mass transfer and high catalyst conversion rates, is the use of 3D printing. By creating optimized structures that prevent channeling and high pressure drops, it is possible to achieve the [...] Read more.
One way to enhance the flow properties of packed bed reactors, including efficient mass transfer and high catalyst conversion rates, is the use of 3D printing. By creating optimized structures that prevent channeling and high pressure drops, it is possible to achieve the desired target. Nevertheless, additively manufactured structures most often possess a limited surface-area-to-volume-ratio, especially as porous printed structures are not standardized yet. One way to achieve surface-enhanced 3D-printed structures is surface modification to introduce surface-initiated polymers. In addition, when stimuli-sensitive polymers are chosen, autonomous process control is prospective. The current publication deals with the application of surface-induced polymerization on 3D-printed structures with the subsequent application as an enzyme carrier. Surface-induced polymerization can easily increase the number of enzymes by a factor of six compared to the non-modified 3D-printed structure. In addition, the swelling behavior of polyacrylic acid is proven, even with immobilized enzymes, enabling smart reaction control. The maximum activity of Esterase 2 (Est2) from Alicyclobacillus acidocaldarius per g carrier, determined after 2 h of polymer synthesis, is 0.61 U/gsupport. Furthermore, universal applicability is shown in aqueous and organic systems, applying an Est2 and Candida antarctica lipase B (CalB) catalyzed reaction and leaving space for improvement due to compatibility of the functionalization process and the here chosen organic solvent. Overall, no enzyme leaching is detectable, and process stability for at least five subsequent batches is ensured. Full article
(This article belongs to the Special Issue Immobilized Biocatalysts II)
Show Figures

Figure 1

Figure 1
<p>Stepwise functionalization process of the support structure made out of PA12 (Polyamide 12). First hydrolysis reaction leads to carboxylic acid and amine groups on the surface (not shown). The carboxylic acid, or rather the hydroxy group, is further utilized for subsequent functionalization steps. Successive aminosilane functionalization leads to surface-standing amine-groups, which are further reacting with the bromine-containing initiator. The initiator works as starting point for the polymerization to achieve polymers, covalently bound to the surface. After PtBA (poly(<span class="html-italic">tert</span>-butyl acrylate)) synthesis, acid hydrolysis is applied to convert PtBA (blue) to the stimuli-sensitive PAA (polyacrylic acid; green), where enzymes are attached via covalent EDC/NHS coupling. Abbreviations: 3-aminopropyltrimethoxysilane (APTMS), 2-bromoisobutyryl bromide (BiBB), tetrahydrofuran (THF), triethylamine (TEA), N,N,N′,N″N″-pentamethyldiethylenetriamine (PMDETA), 2-bromoisobutyrate (EBiB), tin-(II) 2-ethylhexanoate (Sn(II)EH), <span class="html-italic">tert</span>-butyl acrylate (tBA), 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide hydrochloride (EDC), sulfo-<span class="html-italic">N</span>-hydroxysuccinimide (NHS).</p>
Full article ">Figure 2
<p>X-ray photoelectron spectroscopy (XPS) wide-scan of (<b>A</b>) PA12, (<b>B</b>) APTMS functionalized PA12, (<b>C</b>) BiBB functionalized PA12, (<b>D</b>) PtBA-grafted PA12.</p>
Full article ">Figure 3
<p>Fourier-transform infrared spectroscopy (FTIR) analysis of (<b>A</b>) PA12 and acid hydrolyzed PA12, (<b>B</b>) APTMS and BiBB functionalized PA12, (<b>C</b>) PtBA-grafted 3D-printed PA12 in comparison to PA12 and PtBA. The presented range of wavenumbers is chosen depending on the most significant changes within the spectra.</p>
Full article ">Figure 4
<p>Scanning electron microscopy (SEM) images of (<b>A</b>) PA12 3D-printed structure and (<b>B</b>) PtBA-grafted PA12 3D-printed structure. SEM images were captured with a secondary electron detector.</p>
Full article ">Figure 5
<p>Fluorescence microscopy images of GFP (green fluorescent protein) immobilized on (<b>A</b>) PA12 3D-printed structure and (<b>B</b>) PAA-grafted PA12 3D-printed structure.</p>
Full article ">Figure 6
<p>Activity of CalB (<span class="html-italic">Candida antarctica</span> lipase B) without additional ions (control) and with 1 mM of SnCl<sub>2</sub> (<b>A</b>), activity of Est2 without additional ions (control) and with 1 mM of SnCl<sub>2</sub> (<b>B</b>). Both enzymes were screened according to the standard activity assay in triplicates.</p>
Full article ">Figure 7
<p>Dynamic light scattering (DLS) measurement of PA12 particles with grafted polymer and immobilized enzymes (Est2). Polymer synthesis time varies between unmodified (0 h) and 6 h. An increasing median diameter is visible with increasing synthesis time. Additionally, all particles, besides unmodified ones, proof significant swelling behavior. All measurements were conducted in triplicate.</p>
Full article ">Figure 8
<p>PAA-grafted PA12 particles were prepared with different synthesis time, thus different PAA layer thickness. Subsequently, enzymes (Est2) were immobilized and tested for their activity at 40 °C with the standard activity assay (5 mM <span class="html-italic">p</span>NP-acetate, 10% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) acetonitrile in 163 mM PBS buffer (pH 7.0); triplicates). Increasing layer thickness up to 2 h synthesis time leads to higher enzyme activity per g<sub>support</sub>. However, longer synthesis times led to lower enzyme activity per mass, hinting to diffusion limitation.</p>
Full article ">Figure 9
<p>Activity of immobilized enzyme (Est2) per g<sub>support</sub> in dependency of the incubation time in 163 mM PBS buffer, pH 7.0. The increase in activity is constant over 24 h, leading to the conclusion that the synthesized PAA-layer has a very slow swelling response, which is to be optimized for future autonomous process control. Standard activity test was performed at 40 °C, 5 mM <span class="html-italic">p</span>NP-acetate in 10% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) acetonitrile in 163 mM PBS buffer.</p>
Full article ">Figure 10
<p>Comparison of CalB immobilized on PA12-PAA (PA12-PAA-CalB) and on PA12 (PA12-CalB) 3D-printed structure. Faster conversion was determined for PA12-CalB, even with similar amount of enzyme immobilized compared to PA12-PAA-CalB. Reaction conditions: 30 °C, in 1,4-dioxane, 300 mM cinnamyl alcohol, 600 mM vinyl acetate, 300 rpm.</p>
Full article ">Figure 11
<p>SEM images of PA12-PAA-CalB before (<b>A</b>) and after (<b>B</b>) application in organic solvent (1,4-dioxane) for 60 min at 30 °C. SEM images were captured with a secondary electron detector.</p>
Full article ">Figure 12
<p>Comparison of Est2 immobilized on PA12, and PA12-PAA, as well as non-immobilized Est2 (free Est2). Faster conversion was detected for PAA-grafted PA12 structure; still, highest activity is measured for free Est2. Reaction conditions: 20 mM butyl levulinate in 10% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) acetonitrile and 163 mM PBS buffer (pH 7.0), 40 °C, 300 rpm, respective 3D-printed structure, or 1.92 µg/mL Est2 (0.1 U/mg).</p>
Full article ">Figure 13
<p>Recyclability of PAA-grafted 3D-printed PA12 structure with immobilized Est2 in aqueous media. Product concentration below detection limit after five runs. First and second run proofs showed almost 50% reduction of conversion, with stable results between runs two and five. Reaction conditions: 20 mM butyl levulinate in 10% (<span class="html-italic">v</span>/<span class="html-italic">v</span>) acetonitrile and 163 mM PBS buffer (pH 7.0), 40 °C, 300 rpm, respective 3D-printed structure.</p>
Full article ">
18 pages, 10185 KiB  
Article
Structured Catalyst for Indirect Internal Reforming (IIR) of Biogas in Solid Oxide Fuel Cell (SOFC)
by Anna Prioriello, Leonardo Duranti, Igor Luisetto, Frederick Sanna, Claudio Larosa, Maria Luisa Grilli and Elisabetta Di Bartolomeo
Catalysts 2023, 13(7), 1129; https://doi.org/10.3390/catal13071129 - 20 Jul 2023
Cited by 4 | Viewed by 1946
Abstract
The aim of this work is the development of a structured catalyst for the dry reforming of biogas to be used as a pre–reformer in the indirect internal reforming configuration (IIR) of solid oxide fuel cells (SOFCs). The structured catalyst is based on [...] Read more.
The aim of this work is the development of a structured catalyst for the dry reforming of biogas to be used as a pre–reformer in the indirect internal reforming configuration (IIR) of solid oxide fuel cells (SOFCs). The structured catalyst is based on NiCrAl foams coated with ruthenium (nominal loading 3.0 wt%) supported on a CaZr0.85Sm0.15O3−δ (CZS) perovskite oxide. The powder is produced by solution combustion synthesis and deposited on metallic foams by the wash–coating method. Catalytic tests for the dry reforming of methane (DRM) reaction are carried out at 850 °C, 700 °C and 550 °C for an overall 50 h with CH4/CO2 = 1 and p = 1.3 bar at different gas hourly space velocities (GHSVs). The final goal is a proof–of–concept: a laboratory validation of an IIR–SOFC fed by biogas. The carbon amount on spent structured catalysts is evaluated by thermogravimetric analysis and microstructural/compositional investigation. Full article
(This article belongs to the Special Issue New Trends in Electrocatalysis for CO2 Conversion)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic solution combustion synthesis procedure. (<b>i</b>) glycine is added to the solution of metal cations; then, water is evaporated overnight at 70 °C to form a gel; (<b>ii</b>) temperature is risen to 300 °C to ignite the combustion of the gel and form precursors ashes in the form of a black powder; (<b>iii</b>) the obtained brown oxide powder indicates the final perovskite formation, which is obtained after calcination at 850 °C for 5 h.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD plot of Ru3CZS precursor ashes obtained after combustion, the CaZrO<sub>3</sub> reference JCPDS card is added for comparison; (<b>b</b>) Thermogravimetric analysis profile and DTG curve resulting from Ru3CZS precursors, studied from RT to 1000 °C in flowing air; (<b>c</b>) XRD plot of CZS and Ru3CZS after the 5 h calcination step at 850 °C, CaZrO<sub>3</sub> and RuO<sub>2</sub> reference JCPDS cards are added for comparison; (<b>d</b>) 20° &lt; 2θ &lt; 40° enlargement of (<b>c</b>); (<b>e</b>) FE–SEM micrograph of CZS sample after reduction in 5% H<sub>2</sub>/Ar at 850 °C; (<b>f</b>) FE–SEM micrograph of Ru3CZS sample after reduction in 5% H<sub>2</sub>/Ar at 850 °C.</p>
Full article ">Figure 3
<p>FE–SEM micrographs of: (<b>a</b>) NiCrAl foam (larger magnification in the inset) and (<b>b</b>) NiCrAl foam coated with Ru3CSZ catalyst layer (larger magnification in the inset).</p>
Full article ">Figure 4
<p>H<sub>2</sub>–TPR profile of the magnified (5×) bare NiCrAl foam (a); CZS perovskite oxide (b); and Ru3CZS (c).</p>
Full article ">Figure 5
<p>Schematization of the fixed–bed quartz reactor used for the structured catalysts activity assessment.</p>
Full article ">Figure 6
<p>Reaction condition as a function of temperature at different total flows (50, 100 and 200 cm<sup>3</sup>·min<sup>−1</sup>) with CH<sub>4</sub>/CO<sub>2</sub> = 1 composition: (<b>a</b>) CH<sub>4</sub> (closed symbols) and CO<sub>2</sub> (open symbols) conversion (%); (<b>b</b>) H<sub>2</sub>/CO molar ratio.</p>
Full article ">Figure 7
<p>(<b>a</b>) CH<sub>4</sub> and (<b>b</b>) CO<sub>2</sub> conversion (%) vs. time of stream temperature at different total flows (50, 100 and 200 cm<sup>3</sup>·min<sup>−1</sup>) and CH<sub>4</sub>/CO<sub>2</sub> = 1; the temperature profile (purple line) is reported; (<b>c</b>) H<sub>2</sub>/CO molar ratio vs. time of stream at different total flows (50, 100 and 200 cm<sup>3</sup>·min<sup>−1</sup>) and CH<sub>4</sub>/CO<sub>2</sub> = 1; the temperature profile (purple line) is reported.</p>
Full article ">Figure 8
<p>(<b>a</b>) Scheme of packed reactor with labels, positions of the foams and micrographs of A1, A3, A6, A9, and A11 foams; (<b>b</b>) mass loss (%) of A1, A3, A6, A9, A11 between 25 and 1000 °C.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>) Scheme of packed reactor with labels, positions of the foams and micrographs of A1, A3, A6, A9, and A11 foams; (<b>b</b>) mass loss (%) of A1, A3, A6, A9, A11 between 25 and 1000 °C.</p>
Full article ">Figure 9
<p>SEM micrograph and EDX mapping results of a localized fractured coating in an A11 spent sample.</p>
Full article ">Figure 10
<p>(<b>a</b>) I–V and power density curves of the SOFC apparatus tested at 850 °C in dry hydrogen, with (red) and without (black) the structured catalysts; (<b>b</b>) Nyquist plot with EIS data recorded at open circuit for the cells tested in (<b>a</b>). Data were fitted with an equivalent <span class="html-italic">LR(QR)(QR)</span> circuit, which is depicted in the upper part of the panel.</p>
Full article ">Figure 11
<p>(<b>a</b>) Open circuit voltage vs. time trends upon switching the gas feed from hydrogen to DRM mixture (50% CH<sub>4</sub> + 50% CO<sub>2</sub>) for: the SOFC alone (black), the uncoated foams + SOFC (green), the Ru3CZS–coated foams + SOFC (red); (<b>b</b>) I–V and power density curves of the three systems described in (<b>a</b>) tested at 850 °C in DRM mixture; (<b>c</b>) EIS spectra comparison for the complete IIR–SOFC system tested at open circuit in 100% H<sub>2</sub> (blue) and DRM mixture (red). Data were fitted with an equivalent <span class="html-italic">LR(QR)(QR)</span> circuit, ohmic resistance was set to 0 to ease R<sub>pol</sub> comparison.</p>
Full article ">Figure 11 Cont.
<p>(<b>a</b>) Open circuit voltage vs. time trends upon switching the gas feed from hydrogen to DRM mixture (50% CH<sub>4</sub> + 50% CO<sub>2</sub>) for: the SOFC alone (black), the uncoated foams + SOFC (green), the Ru3CZS–coated foams + SOFC (red); (<b>b</b>) I–V and power density curves of the three systems described in (<b>a</b>) tested at 850 °C in DRM mixture; (<b>c</b>) EIS spectra comparison for the complete IIR–SOFC system tested at open circuit in 100% H<sub>2</sub> (blue) and DRM mixture (red). Data were fitted with an equivalent <span class="html-italic">LR(QR)(QR)</span> circuit, ohmic resistance was set to 0 to ease R<sub>pol</sub> comparison.</p>
Full article ">
16 pages, 4349 KiB  
Article
Benzene Oxidation over Pt Loaded on Fly Ash Zeolite X
by Yuri Kalvachev, Totka Todorova, Hristo Kolev, Daniel Merker and Cyril Popov
Catalysts 2023, 13(7), 1128; https://doi.org/10.3390/catal13071128 - 20 Jul 2023
Cited by 2 | Viewed by 1212
Abstract
In the present study, zeolite X (FANaX) was synthesized from coal fly ash (FA) by a two-step high-temperature method. In order to follow the effect of different contaminants in the starting coal ash, zeolite X was also synthesized from pure chemicals according to [...] Read more.
In the present study, zeolite X (FANaX) was synthesized from coal fly ash (FA) by a two-step high-temperature method. In order to follow the effect of different contaminants in the starting coal ash, zeolite X was also synthesized from pure chemicals according to a classical recipe (NaX). Iron was loaded on this reference zeolite with the amount which was contained in the coal FA. The final catalytic samples were obtained by wet impregnation of Pt nanoparticles on both types of zeolite crystals. The most active samples in the benzene oxidation were the platinum-modified ones and, among them, the Pt-impregnated FA zeolite (Pt FANaX). The comparison of the catalytic activity of Pt FANaX with the reference PtFe NaX zeolite showed a temperature difference of 10 °C in favor of Pt FANaX at 50% benzene conversion. From these results, it can be concluded that FA zeolites are a good, cheaper and environmentally friendly alternative to traditional zeolites, synthesized from pure chemicals, which can be applied in the preparation of catalysts for the purification of gaseous mixtures from harmful organic compounds. Full article
(This article belongs to the Section Environmental Catalysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>A</b>) reference zeolite X obtained from chemicals (NaX) and its metal-impregnated analogs: Pt NaX, Fe NaX, PtFe NaX and (<b>B</b>) FA zeolite X (FANaX) and Pt impregnated FA sample with labeled diffraction peaks characteristic for zeolite X (X) and sodalite (S).</p>
Full article ">Figure 2
<p>SEM images of (<b>A</b>) reference zeolite X obtained from chemicals (NaX), (<b>B</b>) zeolite X from fly ash (FANaX) and (<b>C</b>) starting fly ash.</p>
Full article ">Figure 3
<p>N<sub>2</sub> adsorption/desorption isotherms of the studied catalysts: (<b>A</b>) NaX and Pt NaX, (<b>B</b>) Fe NaX and PtFe NaX, (<b>C</b>) FANaX and Pt FANaX. Insets show the pore size distribution.</p>
Full article ">Figure 4
<p>HRTEM imiges of Pt FeNaX sample (<b>A</b>) and Pt FANaX sample (<b>B</b>).</p>
Full article ">Figure 5
<p>FTIR spectra of reference zeolite X (NaX), zeolite X from FA (FANaX) and their metal-impregnated samples: Fe NaX, Pt NaX, PtFe NaX and Pt FANaX.</p>
Full article ">Figure 6
<p>High-resolution XP Spectra of Al2p/Pt4f (<b>A</b>) and Fe2p (<b>B</b>) core levels in the studied NaX and FANaX samples. Blue line shows Pt4f doublet peak, whereas the green line - overlapping Al2p peak. The open circles represent the experimental data.</p>
Full article ">Figure 7
<p>TPR profiles of reference zeolite X (NaX), zeolite X from FA (FANaX) and their metal-impregnated samples: Fe NaX, Pt NaX, PtFe NaX and Pt FANaX.</p>
Full article ">Figure 8
<p>Temperature dependence of the conversion of benzene on reference zeolite X (NaX), zeolite X from FA (FANaX) and their metal-impregnated samples: Fe NaX, Pt NaX, PtFe NaX and Pt FANaX.</p>
Full article ">
11 pages, 1719 KiB  
Article
Iron-Borophosphate Glass-Catalyzed Regioselective Hydrothiolation of Alkynes under Green Conditions
by Nicoli Catholico, Eduarda A. Tessari, Isis J. A. Granja, Martinho J. A. de Sousa, Jorlandio F. Felix, Flávia Manarin, Marcelo Godoi, Jamal Rafique, Ricardo Schneider, Sumbal Saba and Giancarlo V. Botteselle
Catalysts 2023, 13(7), 1127; https://doi.org/10.3390/catal13071127 - 20 Jul 2023
Viewed by 1242
Abstract
Vinyl sulfides are an important class of organic compounds that have relevant synthetic and biological applications. The best-known approach to realize these compounds is the hydrothiolation of alkynes under different conditions using metals, toxic and carcinogenic solvents. The development of new catalysts using [...] Read more.
Vinyl sulfides are an important class of organic compounds that have relevant synthetic and biological applications. The best-known approach to realize these compounds is the hydrothiolation of alkynes under different conditions using metals, toxic and carcinogenic solvents. The development of new catalysts using materials that are environmentally friendly, low in cost, and easy to handle is highly desirable for this reaction. In this regard, glasses have become an important class of materials, since they can be used as a catalyst for chemical reactions. We prepared and characterized an inexpensive and robust iron-doped borophosphate glass (Fe@NaH2PO4-H3BO3 glass). This eco-friendly material was successfully applied as a catalyst for the hydrothiolation of alkynes under solvent-free conditions, affording the desired vinyl sulfides in good-to-excellent yields, with high stereoselectivity. This method of synthesis is attractive because it enables the reuse of the iron-glass catalyst and the scaling up of reactions. Full article
(This article belongs to the Special Issue Feature Papers in Catalysis in Organic and Polymer Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Powder PXRD analysis for borophosphate glass with (mol) (a) 0% Al<sub>2</sub>O<sub>3</sub> and 0% Fe<sup>3+</sup> ions, (b) 10% Al<sub>2</sub>O<sub>3</sub> and 0% Fe<sup>3+</sup> ions, and (c) 10% Al<sub>2</sub>O<sub>3</sub> and 6% Fe<sup>3+</sup> ions and (d) sample (c) after catalysis.</p>
Full article ">Figure 2
<p>Raman spectrum for borophosphate glass (in mol%) (a) 0% Al<sub>2</sub>O<sub>3</sub> and 0% Fe<sup>3+</sup> ions, (b) 10% Al<sub>2</sub>O<sub>3</sub> and 0% Fe<sup>3+</sup> ions, and (c) 10% Al<sub>2</sub>O<sub>3</sub> and 6% Fe<sup>3+</sup> ions.</p>
Full article ">Figure 3
<p>Synthesis of vinyl sulfides catalyzed by iron-borophosphate glass. Reaction conditions: thiol (0.25 mmol), acetylene (0.25 mmol) and Fe@NaH<sub>2</sub>PO<sub>4</sub>-H<sub>3</sub>BO<sub>3</sub> (10 mg). Isolation of products using column chromatography and stereoisomers determined by <sup>1</sup>H NMR spectroscopy.</p>
Full article ">Figure 4
<p>Scaling up reaction for the synthesis of 1.1 g of compound <b>3a</b>.</p>
Full article ">Figure 5
<p>Recyclability of the catalyst.</p>
Full article ">
13 pages, 1348 KiB  
Article
Atom Transfer Radical Addition via Dual Photoredox/Manganese Catalytic System
by Vladislav S. Kostromitin, Vitalij V. Levin and Alexander D. Dilman
Catalysts 2023, 13(7), 1126; https://doi.org/10.3390/catal13071126 - 19 Jul 2023
Cited by 3 | Viewed by 2053
Abstract
Atom transfer radical addition of bromonitromethane and 1,2-dibromotetrafluoroethane to alkenes is described. The reaction is performed under blue light irradiation using two catalysts: 4CzIPN and manganese (II) bromide. The cyanoarene photocatalyst serves for the redox activation of starting organic bromide, while the manganese [...] Read more.
Atom transfer radical addition of bromonitromethane and 1,2-dibromotetrafluoroethane to alkenes is described. The reaction is performed under blue light irradiation using two catalysts: 4CzIPN and manganese (II) bromide. The cyanoarene photocatalyst serves for the redox activation of starting organic bromide, while the manganese salt facilitates the trapping of the alkyl radical with the formation of the carbon–bromine bond. Full article
(This article belongs to the Special Issue Free Radicals in Catalysis, Organic Synthesis, and Material Science)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Scheme 1
<p>ATRA reaction mechanism.</p>
Full article ">Scheme 2
<p>Synthesis of compounds <b>3</b>. Isolated yields are shown. <sup>1</sup> 60 W LED was used.</p>
Full article ">Scheme 3
<p>Gram-scale synthesis and reactions of <b>3u</b>.</p>
Full article ">Scheme 4
<p>Proposed mechanism.</p>
Full article ">Scheme 5
<p>Radical clock experiment.</p>
Full article ">
23 pages, 5762 KiB  
Article
Hydroisomerisation and Hydrocracking of n-Heptane: Modelling and Optimisation Using a Hybrid Artificial Neural Network–Genetic Algorithm (ANN–GA)
by Bashir Y. Al-Zaidi, Ali Al-Shathr, Amal K. Shehab, Zaidoon M. Shakor, Hasan Sh. Majdi, Adnan A. AbdulRazak and James McGregor
Catalysts 2023, 13(7), 1125; https://doi.org/10.3390/catal13071125 - 19 Jul 2023
Cited by 3 | Viewed by 1836
Abstract
In this paper, the focus is on upgrading the value of naphtha compounds represented by n-heptane (n-C7H16) with zero octane number using a commercial zeolite catalyst consisting of a mixture of 75% HY and 25% HZSM-5 [...] Read more.
In this paper, the focus is on upgrading the value of naphtha compounds represented by n-heptane (n-C7H16) with zero octane number using a commercial zeolite catalyst consisting of a mixture of 75% HY and 25% HZSM-5 loaded with different amounts, 0.25 to 1 wt.%, of platinum metal. Hydrocracking and hydroisomerisation processes are experimentally and theoretically studied in the temperature range of 300–400 °C and under various contact times. A feedforward artificial neural network (FFANN) based on two hidden layers was used for the purpose of process modelling. A total of 80% of the experimental results was used to train the artificial neural network, with the remaining results being used for evaluation and testing of the network. Tan-sigmoid and log-sigmoid transfer functions were used in the first and second hidden layers, respectively. The optimum number of neurons in hidden layers was determined depending on minimising the mean absolute error (MAE). The best ANN model, represented by the multilayer FFANN, had a 4–24–24–12 topology. The ANN model accurately simulates the process in which the correlation coefficient (R2) was found to be 0.9918, 0.9492, and 0.9426 for training, validation, and testing, respectively, and an average of 0.9767 for all data. In addition, the operating conditions of the process were optimised using the genetic algorithm (GA) towards increasing the octane number of the products. MATLAB® Version 2020a was utilised to complete all required computations and predictions. Optimal operating conditions were found through the theoretical study: 0.85 wt.% Pt-metal loaded, 359.36 °C, 6.562 H2/n-heptane feed ratio, and 3.409 h−1 weight-hourly space velocity (WHSV), through which the maximum octane number (RON) of 106.84 was obtained. Finally, those operating conditions largely matched what was calculated from the results of the experimental study, where the highest percentage of the resulting isomers was found with about 78.7 mol% on the surface of the catalyst loaded with 0.75 wt.% Pt-metal at 350 °C using a feed ratio of 6.5 H2/n-C7 and WHSV of 2.98 h−1. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>WHSV versus selectivity towards isomer production using different Pt-metal loading at (<b>a</b>) 300 °C, (<b>b</b>) 350 °C, and (<b>c</b>) 400 °C.</p>
Full article ">Figure 2
<p>H<sub>2</sub>/<span class="html-italic">n</span>-C<sub>7</sub> feed ratio versus selectivity towards isomer production using different Pt-metal loading at (<b>a</b>) 300 °C, (<b>b</b>) 350 °C, and (<b>c</b>) 400 °C.</p>
Full article ">Figure 3
<p>Pt-metal loading versus selectivity towards isomer production at 300, 350, and 400 °C using (<b>a</b>) WHSV = 2.98 h<sup>−1</sup> and H<sub>2</sub>/<span class="html-italic">n</span>-C<sub>7</sub> = 6.5, (<b>b</b>) WHSV = 5.03 h<sup>−1</sup> and H<sub>2</sub>/<span class="html-italic">n</span>-C<sub>7</sub> = 8.5, and (<b>c</b>) WHSV = 7.61 h<sup>−1</sup> and H<sub>2</sub>/<span class="html-italic">n</span>-C<sub>7</sub> = 10.5.</p>
Full article ">Figure 4
<p>Mean absolute error achieved for different numbers of hidden layers and neurons: (<b>a</b>) one hidden layer and (<b>b</b>) two hidden layers of equal neurons.</p>
Full article ">Figure 5
<p>Optimum selected topology for ANN network.</p>
Full article ">Figure 6
<p>SSE predictions regarding the number of epochs.</p>
Full article ">Figure 7
<p>Predicted results versus experimental data.</p>
Full article ">Figure 8
<p>Comparison between the real and predicted mole fraction of product components.</p>
Full article ">Figure 9
<p>ANN–GA simulation–optimisation procedure flowchart.</p>
Full article ">Figure 10
<p>Optimisation of <span class="html-italic">RON</span> with the iteration number.</p>
Full article ">Figure 10 Cont.
<p>Optimisation of <span class="html-italic">RON</span> with the iteration number.</p>
Full article ">Figure 11
<p>Experimental hydroisomerisation/hydrocracking laboratory rig.</p>
Full article ">Figure 12
<p>Artificial neural network structure.</p>
Full article ">
26 pages, 5985 KiB  
Review
Solution Plasma for Surface Design of Advanced Photocatalysts
by Rui Wang, Changhua Wang, Yanmei Xing and Xintong Zhang
Catalysts 2023, 13(7), 1124; https://doi.org/10.3390/catal13071124 - 19 Jul 2023
Cited by 2 | Viewed by 1699
Abstract
Rational design of the surface of photocatalysts can conveniently modulate the photo-stimulated charge separation, influence the surface reaction kinetics, and other pivotal factors in the photocatalytic processes for efficient photocatalysis. Solution plasma, holding promise for mild modification of the surface structure of materials, [...] Read more.
Rational design of the surface of photocatalysts can conveniently modulate the photo-stimulated charge separation, influence the surface reaction kinetics, and other pivotal factors in the photocatalytic processes for efficient photocatalysis. Solution plasma, holding promise for mild modification of the surface structure of materials, has recently been recognized as an emerging technology for surface engineering of high-performance photocatalysts. In this review, we will briefly introduce the fundamentals of solution plasma and its applications in materials preparation and summarize the recent research progress in the surface design of advanced photocatalysts by solution plasma. Lastly, we will indicate some possible new directions. This review is expected to provide an instructive guideline for the surface design of heterogeneous photocatalysts by solution plasma. Full article
(This article belongs to the Special Issue Surface Microstructure Design for Advanced Catalysts)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of plasma discharging in the liquid phase form (<b>a</b>) rod-rod, (<b>b</b>) needle-plate, (<b>c</b>) wire-plate, and (<b>d</b>) plate-hole (in an insulation plate)-plate types. Schematic diagram of gas-liquid plasma discharge from (<b>e</b>) needle-plate, (<b>f</b>) multi-needle-plate, and (<b>g</b>) plate-plate types.</p>
Full article ">Figure 2
<p>Schematic diagram of (<b>a</b>) the SP formation process and (<b>b</b>) the distribution of reaction sites and reactive species in SP.</p>
Full article ">Figure 3
<p>(<b>a</b>) Emission spectra of plasma discharging in water with N<sub>2</sub> or O<sub>2</sub>; (<b>b</b>) the generation of H<sub>2</sub>O<sub>2</sub> in SP with N<sub>2</sub> or O<sub>2</sub>. Adapted with permission from Ref. [<a href="#B67-catalysts-13-01124" class="html-bibr">67</a>]. Copyright © 2023 Elsevier B.V.</p>
Full article ">Figure 4
<p>(<b>a</b>) High-resolution transmission electron microscope images (HRTEM); (<b>b</b>) solid <sup>1</sup>H nuclear magnetic resonance (<sup>1</sup>H NMR) spectra; (<b>c</b>) ultraviolet-visible absorption spectra of TiO<sub>2</sub>-AB and SP-treated TiO<sub>2</sub>-AB samples. (<b>d</b>) from left to right, RDB-PAS spectra of SP-treated TiO<sub>2</sub>-AB, TiO<sub>2</sub>-AB, and the difference value of electron trap density between SP-treated TiO<sub>2</sub>-AB and TiO<sub>2</sub>-AB. (<b>e</b>) Schematic illustration of the electron–hole separation mechanism for TiO<sub>2</sub>-AB and SP-treated TiO<sub>2</sub>-AB during photocatalysis. (<b>f</b>) Photo- and photothermal CO<sub>2</sub> conversion rates over TiO<sub>2</sub>-AB and SP-treated TiO<sub>2</sub>-AB samples under solar light irradiation of 100 mW/cm<sup>2</sup>. Adapted with permission from Ref. [<a href="#B140-catalysts-13-01124" class="html-bibr">140</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) <sup>1</sup>H NMR spectra; (<b>b</b>) RDB-PAS spectra; and (<b>c</b>) a schematic diagram of photoexcited electron transfer of N-TiO<sub>2</sub>, N<sub>2</sub>-bubbled SP-treated N-TiO<sub>2</sub>, and O<sub>2</sub>-bubbled SP-treated N-TiO<sub>2</sub>. Adapted with permission from Ref. [<a href="#B67-catalysts-13-01124" class="html-bibr">67</a>]. Copyright © 2023 Elsevier B.V.</p>
Full article ">Figure 6
<p>(<b>a</b>) From left to right, scanning electron microscopy (SEM) images of the Pt-loaded BiVO<sub>4</sub>, the MnO<sub>x</sub>-loaded BiVO<sub>4</sub>, and the Pt/MnO<sub>x</sub>-loaded BiVO<sub>4</sub>. (<b>b</b>) From left to right, SEM images of the Pt-loaded SP-treated BiVO<sub>4</sub>, the MnO<sub>x</sub>-loaded SP-treated BiVO<sub>4</sub>, and the Pt/MnO<sub>x</sub>-loaded SP-treated BiVO<sub>4</sub>. (<b>c</b>) Positron annihilation lifetime spectra of BiVO<sub>4</sub> and SP-treated BiVO<sub>4</sub>. (<b>d</b>) ESR signal of SP-treated BiVO<sub>4</sub>. (<b>e</b>) Photocatalytic oxygen evolution rate form water over BiVO<sub>4</sub> and SP-treated BiVO<sub>4</sub>. The electron sacrificial agent was AgNO<sub>3</sub>, and the light source was visible light (λ ≥ 420 nm). Adapted from Ref. [<a href="#B143-catalysts-13-01124" class="html-bibr">143</a>].</p>
Full article ">Figure 7
<p>(<b>a</b>) HRTEM images of alcohol-plasma-treated P25; (<b>b</b>) ESR spectra; (<b>c</b>,<b>d</b>) EELS spectra of Ti-L<sub>2,3</sub>; (<b>e</b>) RDB-PAS spectroscopy; (<b>f</b>) Photocatalytic hydrogen evolution rate of P25 and alcohol-plasma-treated P25. (<b>g</b>) Photocatalytic hydrogen evolution rate and AQE of alcohol-plasma-treated P25 located with 0.3 wt% Pt. Methanol was the sacrificial reagent, and the light sources were a 300 W Xe lamp and a 300W Xe lamp with a 365 nm filter. Adapted with permission from Ref. [<a href="#B70-catalysts-13-01124" class="html-bibr">70</a>]. Copyright © 2022 American Chemical Society.</p>
Full article ">Figure 8
<p>(<b>a</b>) From top to bottom, the SEM image and high-angle annular dark-filed scanning transmission electron microscopy (HAADF-STEM) images of CeO<sub>2</sub>. (<b>b</b>) the HAADF-STEM images of Au<sub>1</sub>/CeO<sub>2</sub>-H(N<sub>2</sub>). (<b>c</b>) the HAADF-STEM images of Au/CeO<sub>2</sub>(O<sub>2</sub>). Inset were the ESR spectra of CeO<sub>2</sub>, Au/CeO<sub>2</sub>(O<sub>2</sub>), and Au<sub>1</sub>/CeO<sub>2</sub>-H(N<sub>2</sub>); (<b>d</b>) Normalized Au L<sub>3</sub>-edge XANES spectra and (<b>e</b>) Fourier transform EXAFS functions of Au<sub>1</sub>/CeO<sub>2</sub>-H. (<b>f</b>) The temperature dependence of CO conversion for Au<sub>1</sub>/CeO<sub>2</sub> under light and in the dark. (<b>g</b>) The increment of Au<sub>1</sub>/CeO<sub>2</sub>-H in CO<sub>2</sub> production rate and corresponding AQE under a 375 nm laser at different light intensities. (<b>h</b>) Mass spectra of products obtained by the oxidation of <sup>13</sup>CO over Au<sub>1</sub>/CeO<sub>2</sub>-H. (<b>i</b>) Schematic diagram of the photo-enhanced MvK mechanism over the Au<sub>1</sub>/CeO<sub>2</sub>. Adapted with permission from Ref. [<a href="#B74-catalysts-13-01124" class="html-bibr">74</a>]. Copyright © 2022 Wiley-VCH GmbH.</p>
Full article ">Figure 9
<p>(<b>a</b>) TEM and (<b>b</b>) HRTEM images of the HNb<sub>3</sub>O<sub>8</sub>/C heterojunction; (<b>c</b>) Raman spectra; (<b>d</b>) and (<b>e</b>) X-ray photoelectron spectroscopy (XPS) spectra of HNb<sub>3</sub>O<sub>8</sub> and HNb<sub>3</sub>O<sub>8</sub>/C heterojunction. (<b>f</b>) Photocatalytic hydrogen evolution rate of HNb<sub>3</sub>O<sub>8</sub> and HNb<sub>3</sub>O<sub>8</sub>/C heterojunctions. The light source was a 300 W Xe lamp. Adapted with permission from Ref. [<a href="#B154-catalysts-13-01124" class="html-bibr">154</a>]. Copyright © 2022 American Chemical Society.</p>
Full article ">Figure 10
<p>A concise summary diagram of SP-modified/prepared photocatalysts.</p>
Full article ">
13 pages, 4230 KiB  
Article
Tailoring Morphology in Hydrothermally Synthesized CdS/ZnS Nanocomposites for Extraordinary Photocatalytic H2 Generation via Type-II Heterojunction
by Mianli Huang, Maoqing Yu, Ruiru Si, Xiaojing Zhao, Shuqin Chen, Kewei Liu and Xiaoyang Pan
Catalysts 2023, 13(7), 1123; https://doi.org/10.3390/catal13071123 - 19 Jul 2023
Cited by 5 | Viewed by 1403
Abstract
CdS@ZnS core shell nanocomposites were prepared by a one-pot hydrothermal route. The morphology of the composite was tuned by simply changing the Zn2+ precursor concentration. To characterize the samples prepared, various techniques were employed, including XRD, FESEM, TEM, XPS and UV-vis DRS. [...] Read more.
CdS@ZnS core shell nanocomposites were prepared by a one-pot hydrothermal route. The morphology of the composite was tuned by simply changing the Zn2+ precursor concentration. To characterize the samples prepared, various techniques were employed, including XRD, FESEM, TEM, XPS and UV-vis DRS. The band gaps of CdS and ZnS were measured to be 2.26 and 3.32 eV, respectively. Compared with pure CdS, the CdS@ZnS samples exhibited a slight blue shift, which indicated an increased band gap of 2.29 eV. The CdS@ZnS core shell composites exhibited efficient photocatalytic performance for H2 generation under simulated sunlight illumination in contrast to pure CdS and ZnS. Additionally, an optimized H2 generation rate (14.44 mmol·h−1·g−1cat) was acquired at CdS@ZnS-2, which was approximately 4.6 times greater than that of pure CdS (3.12 mmol·h−1·g−1cat). Moreover, CdS@ZnS heterojunction also showed good photocatalytic stability. The process of charge separation over the photocatalysts was investigated using photoelectrochemical analysis. The findings indicate that the CdS@ZnS nanocomposite has efficient charge separation efficiency. The higher H2 generation activity and stability for CdS@ZnS photocatalysts can be attributed to the intimate interface in the CdS@ZnS core–shell structure, which promoted the light absorption intensity and photoinduced charge separation efficiency. It is expected that this study will offer valuable insights into the development of efficient core shell composite photocatalysts. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of the ZnS, CdS and CdS@ZnS samples.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) CdS, (<b>b</b>) CdS@ZnS-1, (<b>c</b>) CdS@ZnS-2, (<b>d</b>) CdS@ZnS-3, (<b>e</b>) CdS@ZnS-4 and (<b>f</b>) ZnS.</p>
Full article ">Figure 3
<p>(<b>a</b>,<b>b</b>) TEM images and (<b>c</b>) HRTEM image of pure CdS NPs; (<b>d</b>,<b>e</b>) TEM images and (<b>f</b>) HRTEM image of CdS@ZnS-2; elemental maps of (<b>g</b>) S, (<b>h</b>) Zn and (<b>i</b>) Cd.</p>
Full article ">Figure 4
<p>XPS spectra of (<b>a</b>) survey, (<b>b</b>) Zn 2p, (<b>c</b>) Cd 3d and (<b>d</b>) S 2p for CdS@ZnS-2.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV-visible diffuse reflection absorption spectrum of the prepared sample; (<b>b</b>) Corresponding Tauc diagrams.</p>
Full article ">Figure 6
<p>(<b>a</b>) Photocatalytic hydrogen production activity and (<b>b</b>) hydrogen production rate diagram of the prepared sample.</p>
Full article ">Figure 7
<p>Cyclic stability diagram of photocatalytic hydrogen production activity for CdS@ZnS-2 sample under simulated sunlight.</p>
Full article ">Figure 8
<p>(<b>a</b>) XRD of CdS@ZnS-2 before and after the recycling test; (<b>b</b>) SEM of CdS@ZnS-2 after recycling test.</p>
Full article ">Figure 9
<p>EIS Nyquist diagram of CdS and CdS@ZnS catalysts.</p>
Full article ">Figure 10
<p>CdS@ZnS band structure mechanism of composite materials.</p>
Full article ">
15 pages, 4077 KiB  
Article
Lithium–Sodium Fly Ash-Derived Catalyst for the In Situ Partial Deoxygenation of Isochrysis sp. Microalgae Bio-Oil
by Nur Adilah Abd Rahman, Fernando Cardenas-Lizana and Aimaro Sanna
Catalysts 2023, 13(7), 1122; https://doi.org/10.3390/catal13071122 - 19 Jul 2023
Cited by 1 | Viewed by 1033
Abstract
The catalytic potential of Na and LiNa fly ash (FA) obtained through a simple solid-state synthesis was investigated for the pyrolysis of Isochrysis sp. microalgae using a fixed bed reactor at 500 °C. While both LiNa-FA and Na-FA catalysts reduced the bio-oil yield [...] Read more.
The catalytic potential of Na and LiNa fly ash (FA) obtained through a simple solid-state synthesis was investigated for the pyrolysis of Isochrysis sp. microalgae using a fixed bed reactor at 500 °C. While both LiNa-FA and Na-FA catalysts reduced the bio-oil yield and increased char and gas production, LiNa-FA was found to enhance the quality of the resulting bio-oil by decreasing its oxygen content (−25 wt.%), increasing paraffins and olefins and decreasing its acidity. The deoxygenation activity of LiNa-FA was attributed to the presence of weak and mild base sites, which enabled dehydration, decarboxylation, ketonisation, and cracking to form olefins. The bio-oil generated with LiNa-FA contained higher amounts of alkanes, alkenes, and carbonated esters, indicating its capacity to chemisorb and partially desorb CO2 under the studied conditions. These findings suggest that LiNa-FA catalysts could be a cost-effective alternative to acidic zeolites for in situ deoxygenation of microalgae to biofuels. Full article
(This article belongs to the Special Issue Catalysis in Biomass Valorization for Fuel and Chemicals)
Show Figures

Figure 1

Figure 1
<p>Comparison of (<b>a</b>) TGA and (<b>b</b>) DTG profiles for the microalga pyrolysis in presence of LiNa-FA, Na-FA, and Li-LSX (from [<a href="#B10-catalysts-13-01122" class="html-bibr">10</a>]). Heating rate of 100 °C/min from 25 °C to 500 °C. Corresponding peak temperatures are also reported.</p>
Full article ">Figure 2
<p>Dynamic CO<sub>2</sub> adsorption (in 20 mL CO<sub>2</sub>/min) for base site assessment. (<b>a</b>) LiNa-FA and (<b>b</b>) Na-FA. Temperature change at rate of 20 °C/min.</p>
Full article ">Figure 3
<p>FTIR of Na-FA and LiNa-FA.</p>
Full article ">Figure 4
<p>SEM images of (<b>a</b>,<b>b</b>) LiNa-FA and (<b>c</b>,<b>d</b>) Na-FA.</p>
Full article ">Figure 5
<p>Product distribution from catalytic and non-catalytic pyrolysis at 500 °C.</p>
Full article ">Figure 6
<p>Nitrogen distribution in the products.</p>
Full article ">Figure 7
<p>Chemical functionalities in bio-oils from GC-MS.</p>
Full article ">Figure 8
<p>Scheme of catalytic conversion of <span class="html-italic">Isochrysis</span> sp. microalgae in presence of LiNa-FA and Na-FA.</p>
Full article ">
13 pages, 3713 KiB  
Article
Effect of Ultraviolet Illumination on the Fixation of Silver Ions on Zinc Oxide Films and Their Photocatalytic Efficiency
by Dobrina Ivanova, Ralitsa Mladenova, Hristo Kolev and Nina Kaneva
Catalysts 2023, 13(7), 1121; https://doi.org/10.3390/catal13071121 - 18 Jul 2023
Cited by 4 | Viewed by 1698
Abstract
This study focuses on the fabrication and characterization of nanostructured zinc oxide films deposited on glass substrates using sol–gel dip-coating methods. The thin films are functionalized with silver ions at various Ag+ concentrations (10−2, 10−3, 10−4 M) [...] Read more.
This study focuses on the fabrication and characterization of nanostructured zinc oxide films deposited on glass substrates using sol–gel dip-coating methods. The thin films are functionalized with silver ions at various Ag+ concentrations (10−2, 10−3, 10−4 M) through room temperature ion fixation process with and without ultraviolet (UV) illumination. Physicochemical characterization techniques, such as employing Scanning Electron Microscopy with Energy-dispersive X-ray spectroscopy (SEM-EDX), X-ray Diffraction (XRD), X-ray Photoelectron Spectroscopy (XPS), Ultraviolet–Visible Spectroscopy and Electron Paramagnetic Resonance (EPR) techniques. The SEM-EDX and XRD confirmed a characteristic ganglia-like structure with a hexagonal crystalline structure. The photocatalytic performance and available surface area of the pure and Ag films are investigated in the removal of methylene blue dye under UV and visible light illumination and in darkness. It is observed that the photocatalytic activity increases proportionally to the Ag+ ion concentration: ZnO < Ag(10−4 M)/ZnO, < Ag(10−3 M)/ZnO < Ag(10−2 M)/ZnO. Moreover, the catalysts modified under UV illumination during the fixation treatment (Ag-UV/ZnO) exhibited a higher photocatalytic efficiency and degraded the dye in comparison with those without a light source (Ag/ZnO). The experimental results are confirmed using total organic carbon (TOC) analysis. The optimal silver concentration (10−2 M) is established, which shows the highest photocatalytic efficiency (in both cases of ion fixation treatment). The results can be used as a guideline for the development of co-catalyst-functionalized semiconductor photocatalysts. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of ZnO (<b>a</b>) and Ag-UV/ZnO (<b>b</b>) films. EDX spectrum of Ag/ZnO (<b>c</b>).</p>
Full article ">Figure 2
<p>XRD patterns of pure and silver co-catalytic-modified semiconductor films.</p>
Full article ">Figure 3
<p>High-resolution XP spectra of Ag-modified zinc oxide films: (<b>a</b>) Ag3d, (<b>b</b>) Zn2p, and (<b>c</b>) O1s regions.</p>
Full article ">Figure 4
<p>EPR spectra of Ag/ZnO and Ag-UV/ZnO films.</p>
Full article ">Figure 5
<p>(<b>a</b>) UV–Vis absorbance spectra and (<b>b</b>) optical bandgap energies of ZnO, Ag/ZnO and Ag-UV/ZnO films.</p>
Full article ">Figure 6
<p>Decolorization kinetics of methylene blue using ZnO, Ag/ZnO (<b>a</b>) and Ag-UV/ZnO (<b>b</b>). The rate constants from photocatalytic processes of sol–gel films modified with silver ions: 10<sup>−2</sup>, 10<sup>−3</sup>, 10<sup>−4</sup> M (<b>c</b>).</p>
Full article ">Figure 7
<p>Percentage of methylene blue dye degradation using pure and silver ZnO photo-fixed without (<b>a</b>) and with (<b>b</b>) UV illumination.</p>
Full article ">Figure 8
<p>Histograms of TOC conversion after photocatalysis—fourth hour of UV illumination.</p>
Full article ">Figure 9
<p>Concentration spectra of dye using pure and silver (10<sup>−2</sup> M)-modified ZnO films photo-fixed without and with ultraviolet illumination. Photocatalytic tests are carried out in the presence of UV (<b>a</b>), visible light (<b>b</b>) and in darkness (<b>c</b>).</p>
Full article ">
16 pages, 2683 KiB  
Article
Structurally Rigid (8-(Arylimino)-5,6,7-trihydroquinolin-2-yl)-methyl Acetate Cobalt Complex Catalysts for Isoprene Polymerization with High Activity and cis-1,4 Selectivity
by Nighat Yousuf, Yanping Ma, Qaiser Mahmood, Wenjuan Zhang, Ming Liu, Rongyan Yuan and Wen-Hua Sun
Catalysts 2023, 13(7), 1120; https://doi.org/10.3390/catal13071120 - 18 Jul 2023
Cited by 8 | Viewed by 1294
Abstract
A series of cobalt complexes bearing (8-(arylimino)-5,6,7-trihydroquinolin-2-yl)methyl acetate ligand framework were prepared using a one-pot synthesis method. These complexes were then extensively investigated for their catalytic performance in isoprene polymerization. In addition to the complexes being characterized via FT-IR spectrum and elemental analysis, [...] Read more.
A series of cobalt complexes bearing (8-(arylimino)-5,6,7-trihydroquinolin-2-yl)methyl acetate ligand framework were prepared using a one-pot synthesis method. These complexes were then extensively investigated for their catalytic performance in isoprene polymerization. In addition to the complexes being characterized via FT-IR spectrum and elemental analysis, the molecular structure of Co1 and Co5 was determined via X-ray diffraction analysis. The analysis revealed a chloride-bridged centrosymmetric binuclear species in which each cobalt center exhibited a distorted square pyramidal geometry. Among the prepared complexes, Co1 demonstrated the highest catalytic activity of 1.37 × 105 g (mol of Co)−1(h)−1, achieving complete monomer conversion and resultant polyisoprene showed high molecular weight (Mn ≥ 2.6 × 105 g/mol). All of the complexes showed preference for the cis-1,4 configuration ranging from 65% to 72%, while the 3,4 monomer insertion units constituted between 27% and 34% of the polymer structure. Moreover, extensive investigations were conducted to assess the impact of reaction parameters and ligand properties on the catalytic activities and microstructural characteristics of the resulting polymer. Full article
(This article belongs to the Special Issue Metal-Organic Catalyst for High Performance Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Molecular structure of <b>Co1</b> (<b>left</b>) and <b>Co5</b> (<b>right</b>) with thermal ellipsoids shown at 30% probability level. All hydrogen atoms are omitted for clarity.</p>
Full article ">Figure 2
<p><sup>1</sup>H and <sup>13</sup>C NMR spectra of the representative sample of polyisoprene obtained using <b>Co1</b>/AlMe<sub>2</sub>Cl (<a href="#catalysts-13-01120-t003" class="html-table">Table 3</a>, entry 4).</p>
Full article ">Figure 3
<p>Polymerization activity (<b>a</b>), average molecular weight and molecular weight distribution (<b>b</b>), and GPC trace (<b>c</b>) for precatalysts <b>Co1</b>–<b>Co6</b>.</p>
Full article ">Figure 4
<p><sup>1</sup>H and <sup>13</sup>C NMR spectra of the representative sample of polyisoprene obtained using <b>Co6</b>/AlMe<sub>2</sub>Cl (<a href="#catalysts-13-01120-t004" class="html-table">Table 4</a>, entry 6).</p>
Full article ">Figure 5
<p>Possible mode of monomer coordination–insertion and corresponding polyisoprene configuration.</p>
Full article ">Chart 1
<p>Iminopyridine-cobalt complexes previously studied for isoprene polymerization (<b>A</b>–<b>C</b>) along with prepared cobalt complexes (<b>D</b>) in this work.</p>
Full article ">Scheme 1
<p>Synthesis of cobalt (II) complexes and the possible mechanism of the incorporation of the methyl acetate group.</p>
Full article ">
12 pages, 2554 KiB  
Article
Application of a Response Surface Method for the Optimization of the Hydrothermal Synthesis of Magnetic NiCo2O4 Desulfurization Catalytic Powders
by Yinke Zhang, Lu Li, Zihan Shang and Hang Xu
Catalysts 2023, 13(7), 1119; https://doi.org/10.3390/catal13071119 - 18 Jul 2023
Cited by 2 | Viewed by 1135
Abstract
In this study, nickel cobaltate (NiCo2O4) powders are employed as a catalyst in conjunction with persulfate for the development of a catalytic oxidation system to enhance fuel desulfurization. The hydrothermal synthesis conditions of NiCo2O4 powders, which [...] Read more.
In this study, nickel cobaltate (NiCo2O4) powders are employed as a catalyst in conjunction with persulfate for the development of a catalytic oxidation system to enhance fuel desulfurization. The hydrothermal synthesis conditions of NiCo2O4 powders, which significantly influenced the desulfurization efficiency, were optimized using a response surface methodology with a Box–Behnken design. These conditions were ranked in the following order: calcination temperature > hydrothermal temperature > calcination time > hydrothermal time. Through the optimization process, the ideal preparation conditions were determined as follows: a hydrothermal temperature of 143 °C, hydrothermal time of 6.1 h, calcination temperature of 330 °C, and calcination time of 3.7 h. Under these optimized conditions, the predicted desulfurization rate was approximately 85.8%. The experimental results closely matched the prediction, yielding a desulfurization rate of around 84%, with a minimal error of only 2.1%. To characterize the NiCo2O4 powders prepared under the optimal conditions, XRD, SEM, and TEM analyses were conducted. The analysis revealed that the microscopic morphology of NiCo2O4 exhibited a rectangular sheet structure, with an average particle size of 20 nm. Additionally, fan-shaped NiCo2O4 particles were observed as a result of linear and bundle agglomerations. Thus, this work is innovative in its ability to synthesize nano-catalysts using hydrothermal synthesis in a controllable manner and establishing a correlation between the hydrothermal synthesis conditions and catalytic activity. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of predicted and experimental values based on the second-order model.</p>
Full article ">Figure 2
<p>The internally studentized residuals at run number from 1 to 29.</p>
Full article ">Figure 3
<p>Response surface plot of the effect of hydrothermal temperature (<b>A</b>), hydrothermal time (<b>B</b>), calcination temperature (<b>C</b>), and calcination time (<b>D</b>) on the desulfurization rate.</p>
Full article ">Figure 4
<p>Optimal experimental conditions from RSM.</p>
Full article ">Figure 5
<p>XRD (<b>a</b>) and SEM (<b>b</b>) images of NiCo<sub>2</sub>O<sub>4</sub> powders.</p>
Full article ">Figure 6
<p>TEM of NiCo<sub>2</sub>O<sub>4</sub> and the formation mechanism of the powders.</p>
Full article ">Figure 7
<p>Cyclic desulfurization experiment of NiCo<sub>2</sub>O<sub>4</sub> powders.</p>
Full article ">
13 pages, 9616 KiB  
Article
Mitigating Co Metal Particle Agglomeration and Enhancing ORR Catalytic Activity through Nitrogen-Enriched Porous Carbon Derived from Biomass
by Yanling Wu, Qinggao Hou, Fangzhou Li, Yuanhua Sang, Mengyang Hao, Xi Tang, Fangyuan Qiu and Haijun Zhang
Catalysts 2023, 13(7), 1118; https://doi.org/10.3390/catal13071118 - 18 Jul 2023
Cited by 2 | Viewed by 1729
Abstract
Biomass-derived porous carbon has gained significant attention as a cost-effective and sustainable material in non-noble metal carbon-based electrocatalysts for the oxygen reduction reaction (ORR). However, during the preparation of transition metal catalysts based on biomass-derived porous carbon, the agglomeration of transition metal atoms [...] Read more.
Biomass-derived porous carbon has gained significant attention as a cost-effective and sustainable material in non-noble metal carbon-based electrocatalysts for the oxygen reduction reaction (ORR). However, during the preparation of transition metal catalysts based on biomass-derived porous carbon, the agglomeration of transition metal atoms often occurs, leading to a notable decline in catalytic activity. In this study, we present a straightforward synthetic approach for the preparation of nitrogen-enriched soybean-derived porous carbon (Co@SP-C-a) as an electrocatalyst for the ORR. To achieve this, we employed a two-step method. In the first step, a chemical activator (KCl) was utilized to enhance the porosity of the self-doped nitrogen biomass carbon material. In the second step, a constant pressure drop funnel technique was employed to uniformly disperse bimetal cobalt/zinc-based zeolitic imidazolium frameworks (ZIF-L and ZIF-67) containing different metal ions (Zn2+ and Co2+) into the activated biomass carbon material. Subsequent high-temperature calcination of the ZIF-L and ZIF-67@SP-C-a composite precursor yielded the Co@SP-C-a catalyst. The obtained catalyst exhibited remarkable ORR activity in an alkaline solution (Eonset = 0.89 V, E1/2 = 0.83 V, JL = −6.13 mA·cm−2) and exceptional long-term stability. This study presents an effective strategy to prevent the agglomeration of metal nanoparticles when integrating them with biomass-based carbon materials, thus leading to enhanced catalytic performance. Full article
(This article belongs to the Special Issue Morphological Effects on Catalytic Reactions)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The synthetic scheme of Co@SP-C-a catalyst.</p>
Full article ">Figure 2
<p>Representative XRD patterns (<b>a</b>); SEM (<b>b</b>), TEM (inset: particle size distribution of Co@SP-C-a) (<b>c</b>), and lattice fringe HR–TEM (<b>d</b>) images of Co@SP-C-a.</p>
Full article ">Figure 3
<p>Raman spectra (<b>a</b>); the nitrogen adsorption/desorption isotherms (<b>b</b>); high-resolution XPS data for (<b>c</b>) C 1s, (<b>d</b>) Co 2p, (<b>e</b>) N 1s, (<b>f</b>) O 1s core levels of Co@SP-C-a.</p>
Full article ">Figure 4
<p>(<b>a</b>) CV curves obtained for SP-C, SP-C-a, and Co@SP-C-a at 50 mV s<sup>−1</sup> in 0.1 M potassium hydroxide saturated with O<sub>2</sub> (solid lines) and N<sub>2</sub> (dotted lines). (<b>b</b>) ORR LSV curves at 10 mV s<sup>−1</sup> under 1600 rpm recorded for the three SP-C-based catalysts and a 20 wt.% Pt/C catalyst in the same electrolyte saturated with oxygen. of (<b>c</b>) LSV curves of Co@SP-C-a at 10 mV s<sup>−1</sup> under various rotation speeds, with the inset showing K-L plots of Co@SP-C-a at different potentials. (<b>d</b>) Tafel slope and electrochemical double-layer capacitance (C<sub>dl</sub>) comparisons of SP-C, SP-C-a, and Co@SP-C-a.</p>
Full article ">
Previous Issue
Next Issue
Back to TopTop