[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,758)

Search Parameters:
Keywords = redox property

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
27 pages, 8897 KiB  
Review
Engineering of g-C3N4 for Photocatalytic Hydrogen Production: A Review
by Yachao Yan, Qing Meng, Long Tian, Yulong Cai, Yujuan Zhang and Yingzhi Chen
Int. J. Mol. Sci. 2024, 25(16), 8842; https://doi.org/10.3390/ijms25168842 - 14 Aug 2024
Abstract
Graphitic carbon nitride (g-C3N4)-based photocatalysts have garnered significant interest as a promising photocatalyst for hydrogen generation under visible light, to address energy and environmental challenges owing to their favorable electronic structure, affordability, and stability. In spite of [...] Read more.
Graphitic carbon nitride (g-C3N4)-based photocatalysts have garnered significant interest as a promising photocatalyst for hydrogen generation under visible light, to address energy and environmental challenges owing to their favorable electronic structure, affordability, and stability. In spite of that, issues such as high charge carrier recombination rates and low quantum efficiency impede its broader application. To overcome these limitations, structural and morphological modification of the g-C3N4-based photocatalysts is a novel frontline to improve the photocatalytic performance. Therefore, we briefly summarize the current preparation methods of g-C3N4. Importantly, this review highlights recent advancements in crafting high-performance g-C3N4-based photocatalysts, focusing on strategies like elemental doping, nanostructure design, bandgap engineering, and heterostructure construction. Notably, sophisticated doping techniques have propelled hydrogen production rates to a 104-fold increase. Ingenious nanostructure designs have expanded the surface area by a factor of 26, concurrently extending the fluorescence lifetime of charge carriers by 50%. Moreover, the strategic assembly of heterojunctions has not only elevated charge carrier separation efficiency but also preserved formidable redox properties, culminating in a dramatic hundredfold surge in hydrogen generation performance. This work provides a reliable and brief overview of the controlled modification engineering of g-C3N4-based photocatalyst systems, paving the way for more efficient hydrogen production. Full article
(This article belongs to the Section Materials Science)
Show Figures

Figure 1

Figure 1
<p>Photocatalytic H<sub>2</sub> production of semiconductor.</p>
Full article ">Figure 2
<p>Molecular structure of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>: (<b>a</b>) triazine structure (C<sub>3</sub>N<sub>3</sub>); (<b>b</b>) 3-s-triazine structure (C<sub>6</sub>N<sub>7</sub>).</p>
Full article ">Figure 3
<p>Schematic diagram of synthesis strategies <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> (gray, blue, white and red balls denote C, N, H and O atoms, respectively).</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–Visible absorbance spectra of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> samples; band gap analysis-Tauc plot of (<b>b</b>) bulk <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and (<b>c</b>) prepared <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>; (<b>d</b>)time-dependent degradation kinetics; (<b>e</b>) photocatalytic degradation efficiency after 60 min of degradation [<a href="#B45-ijms-25-08842" class="html-bibr">45</a>]; (<b>f</b>) N<sub>2</sub> adsorption–desorption isotherms; (<b>g</b>) pore size distributions (insert: detail structural parameters) [<a href="#B46-ijms-25-08842" class="html-bibr">46</a>]; (<b>h</b>) schematic illustration of the synthesis process of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> by melamine, cyanamide, dicyanamide, urea, and thiourea [<a href="#B47-ijms-25-08842" class="html-bibr">47</a>]; SEM images of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/BiOBr prepared at (<b>i</b>) 150 °C and (<b>j</b>) 160 °C; (<b>k</b>) degradation curves and 60 min degradation rates of samples prepared with different composition ratios [<a href="#B48-ijms-25-08842" class="html-bibr">48</a>]; (<b>l</b>) N<sub>2</sub> adsorption–desorption isotherms of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> [<a href="#B49-ijms-25-08842" class="html-bibr">49</a>]. Adapted with permission.</p>
Full article ">Figure 5
<p>TEM images of (<b>a</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and (<b>b</b>) NCN-x; (<b>c</b>) doping process for the NCN-x with an increase in Ni content; (<b>d</b>) photocatalytic hydrogen evolution of NCN-x [<a href="#B54-ijms-25-08842" class="html-bibr">54</a>] (adapted with permission); (<b>e</b>) TEM images of KCN; (<b>f</b>) band gap illustration of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and KCN; (<b>g</b>) photocatalytic H<sub>2</sub> production rate of KCN [<a href="#B59-ijms-25-08842" class="html-bibr">59</a>] (adapted with permission).</p>
Full article ">Figure 6
<p>(<b>a</b>) The substitution of bridging N in triazine ring of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> with carbon atom; (<b>b</b>) X-ray photoelectron spectroscopy analysis of CNU; (<b>c</b>) Nyquist plots; (<b>d</b>) photocurrent response; and (<b>e</b>) hydrogen evolution of BCN and CNU; (<b>f</b>) stability test of the CNU [<a href="#B85-ijms-25-08842" class="html-bibr">85</a>]; TEM images of (<b>g</b>) bulk <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and (<b>h</b>) O/<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>; (<b>i</b>) UV–vis light absorption spectra (insets indicate the corresponding color) and (<b>j</b>) photocatalytic H<sub>2</sub> evolution corresponding apparent rate constants of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and O/<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> [<a href="#B86-ijms-25-08842" class="html-bibr">86</a>] (adapted with permission) (<b>k</b>) TEM images; (<b>l</b>) TCSPC spectra and (<b>m</b>) Photocatalytic H<sub>2</sub> evolution profiles of N-O-CNNS [<a href="#B89-ijms-25-08842" class="html-bibr">89</a>] (adapted with permission).</p>
Full article ">Figure 7
<p>(<b>a</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> of chiral helical nanorods for photocatalytic water splitting and CO<sub>2</sub> conversion [<a href="#B91-ijms-25-08842" class="html-bibr">91</a>] (adapted with permission); (<b>b</b>) photocatalytic hydrogen evolution schematic of OCN nanotubes; (<b>c</b>) photoluminescence spectra of OCN nanotubes; (<b>d</b>) HER of the bulk CN and OCN samples [<a href="#B92-ijms-25-08842" class="html-bibr">92</a>] (adapted with permission).</p>
Full article ">Figure 8
<p>Typical TEM images: (<b>a</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> (540)-T; (<b>b</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>(560)-T; (<b>c</b>) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>(580)-T; (<b>d</b>) EIS of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>(5X0)-T; (<b>e</b>) photocurrent response of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>(5X0)-T [<a href="#B93-ijms-25-08842" class="html-bibr">93</a>] (adapted with permission); (<b>f</b>) time-resolved fluorescence decay spectra of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> nanosheets: a) bulk <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, and b) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> nanosheets; (<b>g</b>) photocatalytic hydrogen evolution of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> nanosheets: a) bulk <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>, and b) <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> nanosheets [<a href="#B94-ijms-25-08842" class="html-bibr">94</a>] (adapted with permission).</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematic illustration of the preparation of the porous <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> nanosheets CN1; (<b>b</b>) TEM images of CN1; (<b>c</b>) photocatalytic H<sub>2</sub> production rate of CN1 [<a href="#B95-ijms-25-08842" class="html-bibr">95</a>] (adapted with permission) (<b>d</b>) UV–Vis diffuse reflectance spectra of HCN [<a href="#B96-ijms-25-08842" class="html-bibr">96</a>] (adapted with permission); (<b>e</b>) photocatalytic H<sub>2</sub> production rate of CNHS [<a href="#B97-ijms-25-08842" class="html-bibr">97</a>] (adapted with permission).</p>
Full article ">Figure 10
<p>(<b>a</b>) Diagram of the chemical mechanism of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Cu<sub>2</sub>O; (<b>b</b>) photocurrent responses curves of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Cu<sub>2</sub>O; (<b>c</b>) photocatalytic hydrogen evolution of prepared photocatalysts under visible light [<a href="#B100-ijms-25-08842" class="html-bibr">100</a>] (adapted with permission); (<b>d</b>) diagram of the chemical mechanism of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Sn<sub>3</sub>O<sub>4</sub>; (<b>e</b>) ultraviolet absorption spectra of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> and <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Sn<sub>3</sub>O<sub>4</sub>; (<b>f</b>) photocatalytic hydrogen evolution of prepared photocatalysts under visible light [<a href="#B101-ijms-25-08842" class="html-bibr">101</a>] (adapted with permission); (<b>g</b>) synthetic route of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Au; (<b>h</b>) Transmission Electron Microscope of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Au; (<b>i</b>) photocatalytic hydrogen evolution of <span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub>/Au [<a href="#B102-ijms-25-08842" class="html-bibr">102</a>] (adapted with permission); (<b>j</b>) HR-TEM images; (<b>k</b>) photoluminescence spectra; and (<b>l</b>) photocatalytic H<sub>2</sub> production rate of Co (Mo-Mo<sub>2</sub>C)/<span class="html-italic">g</span>-C<sub>3</sub>N<sub>4</sub> [<a href="#B108-ijms-25-08842" class="html-bibr">108</a>] (adapted with permission).</p>
Full article ">
16 pages, 4159 KiB  
Article
Effect of Substituted Pyridine Co-Ligands and (Diacetoxyiodo)benzene Oxidants on the Fe(III)-OIPh-Mediated Triphenylmethane Hydroxylation Reaction
by Patrik Török and József Kaizer
Molecules 2024, 29(16), 3842; https://doi.org/10.3390/molecules29163842 - 13 Aug 2024
Viewed by 249
Abstract
Iodosilarene derivatives (PhIO, PhI(OAc)2) constitute an important class of oxygen atom transfer reagents in organic synthesis and are often used together with iron-based catalysts. Since the factors controlling the ability of iron centers to catalyze alkane hydroxylation are not yet fully [...] Read more.
Iodosilarene derivatives (PhIO, PhI(OAc)2) constitute an important class of oxygen atom transfer reagents in organic synthesis and are often used together with iron-based catalysts. Since the factors controlling the ability of iron centers to catalyze alkane hydroxylation are not yet fully understood, the aim of this report is to develop bioinspired non-heme iron catalysts in combination with PhI(OAc)2, which are suitable for performing C-H activation. Overall, this study provides insight into the iron-based ([FeII(PBI)3(CF3SO3)2] (1), where PBI = 2-(2-pyridyl)benzimidazole) catalytic and stoichiometric hydroxylation of triphenylmethane using PhI(OAc)2, highlighting the importance of reaction conditions including the effect of the co-ligands (para-substituted pyridines) and oxidants (para-substituted iodosylbenzene diacetates) on product yields and reaction kinetics. A number of mechanistic studies have been carried out on the mechanism of triphenylmethane hydroxylation, including C-H activation, supporting the reactive intermediate, and investigating the effects of equatorial co-ligands and coordinated oxidants. Strong evidence for the electrophilic nature of the reaction was observed based on competitive experiments, which included a Hammett correlation between the relative reaction rate (logkrel) and the σp (4R-Py and 4R’-PhI(OAc)2) parameters in both stoichiometric (ρ = +0.87 and +0.92) and catalytic (ρ = +0.97 and +0.77) reactions. The presence of [(PBI)2(4R-Py)FeIIIOIPh-4R’]3+ intermediates, as well as the effect of co-ligands and coordinated oxidants, was supported by their spectral (UV–visible) and redox properties. It has been proven that the electrophilic nature of iron(III)-iodozilarene complexes is crucial in the oxidation reaction of triphenylmethane. The hydroxylation rates showed a linear correlation with the FeIII/FeII redox potentials (in the range of −350 mV and −524 mV), which suggests that the Lewis acidity and redox properties of the metal centers greatly influence the reactivity of the reactive intermediates. Full article
(This article belongs to the Special Issue Inorganic Chemistry in Europe)
Show Figures

Figure 1

Figure 1
<p>[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub>-catalyzed hydroxylation of triphenylmethane with PhI(OAc)<sub>2</sub> in the absence and presence of <span class="html-italic">para</span>-substituted pyridines in acetonitrile at 323 K: [Fe<sup>II</sup>(OTf)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−1</sup> M, and [Ph<sub>3</sub>CH]<sub>0</sub> = 3 × 10<sup>−1</sup> M, [pyridine]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 2
<p>[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub>-catalyzed hydroxylation of triphenylmethane with PhI(OAc)<sub>2</sub> in the presence of <span class="html-italic">para</span>-substituted pyridines in acetonitrile at 323 K: (<b>a</b>) the calculated conversion (=TON) values for <span class="html-italic">para</span>-substituted pyridines. (<b>b</b>) Hammett plot of log<span class="html-italic">k</span><sub>rel</sub> against the <span class="html-italic">σ</span><sub>p</sub> of <span class="html-italic">para</span>-substituted pyridines. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−1</sup> M, [Ph<sub>3</sub>CH]<sub>0</sub> = 3 × 10<sup>−1</sup> M, [pyridine]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 3
<p>(<b>a</b>) Formation of [(PBI)<sub>2</sub>(MeCN)Fe<sup>III</sup>(4R-PhIO)]<sup>3+</sup> intermediates in the in situ reaction of <b>1</b> with 4R’-Ph(IOAc)<sub>2</sub> in acetonitrile at 293 K monitored at 760 nm. (<b>b</b>) Hammett plot of log<span class="html-italic">k</span><sub>rel</sub> against the <span class="html-italic">σ</span><sub>p</sub> of <span class="html-italic">para</span>-substituted 4R’-Ph(IOAc)<sub>2</sub> oxidants. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [4R’-PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 4
<p>[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub>-catalyzed hydroxylation of triphenylmethane with 4R’-PhI(OAc)<sub>2</sub> in acetonitrile at 323 K: (<b>a</b>) the calculated conversion (=TON) values for 4R’-PhI(OAc)<sub>2</sub>. (<b>b</b>) Hammett plot of log<span class="html-italic">k</span><sub>rel</sub> against the <span class="html-italic">σ</span><sub>p</sub> of 4R-PhI(OAc)<sub>2</sub>. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [4R’-PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−1</sup> M, and [Ph<sub>3</sub>CH]<sub>0</sub> = 3 × 10<sup>−1</sup> M.</p>
Full article ">Figure 5
<p>Redox properties of the [(PBI)<sub>2</sub>Fe<sup>III</sup>(OIPh)(4R-Py)] intermediates generated in situ by the reaction of <b>1</b> with Ph(IOAc)<sub>2</sub> in acetonitrile at 293 K. (<b>a</b>) Cyclic voltammograms of [(PBI)<sub>2</sub>Fe<sup>III</sup>(OIPh)(4R-Py)] intermediates. (<b>b</b>) <span class="html-italic">E</span><sub>1/2</sub> vs. σ values for [(PBI)<sub>2</sub>[(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates. Conditions: [[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub>= 2 × 10<sup>−3</sup> M, [4R-Py]<sub>0</sub> = 1 × 10<sup>−2</sup> M, in (0.1 M TBAClO<sub>4</sub>) CH<sub>3</sub>CN (10 cm<sup>3</sup>), scan rate: 1500 mV/s.</p>
Full article ">Figure 6
<p>Decrease in absorbance of [(PBI)<sub>2</sub>[(4Ac-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediate in the stoichiometric oxidation of triphenylmethane at 293 K in acetonitrile. Inset: time course of the decay monitored at 745 nm. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1.2 × 10<sup>−3</sup> M, [triphenylmethane]<sub>0</sub> = 1 × 10<sup>−1</sup> M, [4Ac-Py]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 7
<p>Stoichiometric hydroxylation of triphenylmethane with [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates generated in situ by the reaction of <b>1</b> with Ph(IOAc)<sub>2</sub> and 4R-Py derivatives in acetonitrile at 293 K: (<b>a</b>) monitoring the decrease in absorbance of [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates by UV–Vis spectroscopy at 723–760 nm over time at 293 K. (<b>b</b>) Hammett plot of log<span class="html-italic">k</span><sub>rel</sub> against the <span class="html-italic">σ</span><sub>p</sub> of 4R-Py. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1.2 × 10<sup>−3</sup> M, [Ph<sub>3</sub>CH]<sub>0</sub> = 1 × 10<sup>−1</sup> M, [4R-Py]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 8
<p>Stoichiometric hydroxylation of triphenylmethane with [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates generated in situ by the reaction of [Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub> with Ph(IOAc)<sub>2</sub> and 4R-Py derivatives in acetonitrile at 293 K: (<b>a</b>) log(<span class="html-italic">k</span><sub>rel</sub>) against <span class="html-italic">E</span><sub>1/2</sub> for [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates. (<b>b</b>) log(<span class="html-italic">k</span><sub>rel</sub>) against ν for [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> intermediates. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1.2 × 10<sup>−3</sup> M, [Ph<sub>3</sub>CH]<sub>0</sub> = 1 × 10<sup>−1</sup> M, [4R-Py]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 9
<p>UV–Vis spectral properties of [(PBI)<sub>2</sub>[(X)Fe<sup>III</sup>(OIPh)]<sup>3+</sup> (X = CH<sub>3</sub>CN, Py, PyO) intermediates at 293 K in acetonitrile. [<b>1</b>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1.2 × 10<sup>−3</sup> M, [triphenylmethane]<sub>0</sub> = 1 × 10<sup>−1</sup> M, [X]<sub>0</sub> = 1 × 10<sup>−2</sup> M.</p>
Full article ">Figure 10
<p>Stoichiometric hydroxylation of triphenylmethane with [(PBI)<sub>2</sub>(MeCN)Fe<sup>III</sup>(OIPh-4R)]<sup>3+</sup> (R = -Cl, -H, -Me, -OMe) intermediates generated in situ by the reaction of <b>1</b> with 4R-Ph(IOAc)<sub>2</sub> (R = -Cl, -H, -Me, -OMe) oxidants in acetonitrile at 293 K: (<b>a</b>) Monitoring the decrease in absorbance of [(PBI)<sub>2</sub>(CH<sub>3</sub>CN)Fe<sup>III</sup>(OIPh-4R)]<sup>3+</sup> intermediates by UV–Vis spectroscopy at 760 nm over time at 293 K. (<b>b</b>) Hammett plot of log<span class="html-italic">k</span><sub>rel</sub> against the <span class="html-italic">σ</span><sub>p</sub> of [(PBI)<sub>2</sub>(CH<sub>3</sub>CN)Fe<sup>III</sup>(OIPh-4R)]<sup>3+</sup>intermediates. [[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub>]<sub>0</sub> = 1 × 10<sup>−3</sup> M, [4R-PhI(OAc)<sub>2</sub>]<sub>0</sub> = 1.2 × 10<sup>−3</sup> M, [Ph<sub>3</sub>CH]<sub>0</sub> = 1 × 10<sup>−1</sup> M.</p>
Full article ">Scheme 1
<p>[Fe<sup>II</sup>(PBI)<sub>3</sub>](OTf)<sub>2</sub> (<b>1</b>) catalyzed oxidation of triphenylmethane mediated by [(PBI)<sub>2</sub>(4R-Py)Fe<sup>III</sup>(4R’-PhIO)]<sup>3</sup>+ intermediates in the presence of 4R’-PhI(OAc)<sub>2</sub> terminal oxidants and 4R-Py co-ligands.</p>
Full article ">Scheme 2
<p>Proposed mechanistic pathways for the iron(III)–iodosylarene ([(PBI)<sub>2</sub>[(4R-Py)Fe<sup>III</sup>(OIPh-4R)]<sup>3+</sup>) mediated oxidation of triphenylmethane.</p>
Full article ">
18 pages, 4088 KiB  
Article
Enhancing the Performance of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) Perovskites as Catalysts for CO Oxidation by Decreasing the Ba Content
by Á. Díaz-Verde and M. J. Illán-Gómez
Nanomaterials 2024, 14(16), 1334; https://doi.org/10.3390/nano14161334 - 10 Aug 2024
Viewed by 292
Abstract
Mixed oxides featuring perovskite-type structures (ABO3) offer promising catalytic properties for applications focused on the control of atmospheric pollution. In this work, a series of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) samples have been synthesized, characterized [...] Read more.
Mixed oxides featuring perovskite-type structures (ABO3) offer promising catalytic properties for applications focused on the control of atmospheric pollution. In this work, a series of BaxMnO3 (x = 1, 0.9, 0.8 and 0.7) samples have been synthesized, characterized and tested as catalysts for CO oxidation reaction in conditions close to that found in the exhausts of last-generation automotive internal combustion engines. All samples were observed to be active as catalysts for CO oxidation during CO-TPRe tests, with Ba0.7MnO3 (B0.7M) being the most active one, as it presents the highest amount of oxygen vacancies (which act as active sites for CO oxidation) and Mn (IV), which features the highest levels of reducibility and the best redox properties. B0.7M has also showcased a high stability during reactions at 300 °C, even though a slightly lower CO conversion is achieved during the second consecutive reaction cycle. This performance appears to be related to the decrease in the Mn (IV)/Mn (III) ratio. Full article
(This article belongs to the Section Nanophotonics Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of BxM samples.</p>
Full article ">Figure 2
<p>XPS spectra of the Mn 2p<sup>3/2</sup> (<b>a</b>), Mn 3p (<b>b</b>), Ba 3d<sup>5/2</sup> (<b>c</b>) and O 1s (<b>d</b>) transitions.</p>
Full article ">Figure 3
<p>H<sub>2</sub>-TPR consumption profiles for BxM samples, for Mn<sub>2</sub>O<sub>3</sub> and MnO<sub>2</sub> references (<b>a</b>), and for H<sub>2</sub> consumption (mL H<sub>2</sub> (g of cat)<sup>−1</sup>) (<b>b</b>).</p>
Full article ">Figure 4
<p>CO conversion profiles of the first (solid lines) and the second cycle (dotted lines) of CO-TPR tests for BxM samples.</p>
Full article ">Figure 5
<p>O<sub>2</sub>-TPD profiles for BxM samples.</p>
Full article ">Figure 6
<p>CO conversion profiles for BxM and 1% Pt/Al<sub>2</sub>O<sub>3</sub> samples in the 0.1% CO/1% O<sub>2</sub>/He (<b>a</b>); 1% CO/1% O<sub>2</sub>/He (<b>b</b>); and 1% CO/10% O<sub>2</sub>/He (<b>c</b>) reactant mixtures.</p>
Full article ">Figure 7
<p>CO conversion profiles of B0.7M at 300 °C in the 1% CO/He reactant mixture.</p>
Full article ">Figure 8
<p>XPS spectra of the Mn 2p<sub>3/2</sub> (<b>a</b>) and O 1s (<b>b</b>) transitions for the fresh and the spent B0.7M samples.</p>
Full article ">
17 pages, 3982 KiB  
Review
Nicotinamide Riboside: What It Takes to Incorporate It into RNA
by Felix Wenzek, Alexander Biallas and Sabine Müller
Molecules 2024, 29(16), 3788; https://doi.org/10.3390/molecules29163788 - 10 Aug 2024
Viewed by 205
Abstract
Nicotinamide is an important functional compound and, in the form of nicotinamide adenine dinucleotide (NAD), is used as a co-factor by protein-based enzymes to catalyze redox reactions. In the context of the RNA world hypothesis, it is therefore reasonable to assume that ancestral [...] Read more.
Nicotinamide is an important functional compound and, in the form of nicotinamide adenine dinucleotide (NAD), is used as a co-factor by protein-based enzymes to catalyze redox reactions. In the context of the RNA world hypothesis, it is therefore reasonable to assume that ancestral ribozymes could have used co-factors such as NAD or its simpler analog nicotinamide riboside (NAR) to catalyze redox reactions. The only described example of such an engineered ribozyme uses a nicotinamide moiety bound to the ribozyme through non-covalent interactions. Covalent attachment of NAR to RNA could be advantageous, but the demonstration of such scenarios to date has suffered from the chemical instability of both NAR and its reduced form, NARH, making their use in oligonucleotide synthesis less straightforward. Here, we review the literature describing the chemical properties of the oxidized and reduced species of NAR, their synthesis, and previous attempts to incorporate either species into RNA. We discuss how to overcome the stability problem and succeed in generating RNA structures incorporating NAR. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Scheme of the redox cycle of an artificial nicotinamide-containing RNA structure with oxidoreductase activity.</p>
Full article ">Figure 2
<p>Nucleophilic degradation of NAR by either attack onto the heterocycle or the anomeric carbon.</p>
Full article ">Figure 3
<p>Selection of key nicotinamide-based compounds important for incorporation into RNA structures.</p>
Full article ">Figure 4
<p>Acid-catalyzed vinylamine addition, anomeric displacement, and isomerization leading to degradation of NARH.</p>
Full article ">Figure 5
<p>Alteration in resonance contribution of the carbonyl substituent of NARH affecting the rate of initial protonation. The stabilization induced by an additional resonance structure facilitates protonation of the nicotinamide moiety.</p>
Full article ">Figure 6
<p>Access to NAR(H) derivatives by (<b>a</b>) nucleophilic substitution of halo sugars; (<b>b</b>) Lewis acid catalyzed glycosylation of per-acetylated ribose, and (<b>c</b>) formation of the nicotinamide moiety through Zincke reaction.</p>
Full article ">Figure 7
<p>Synthetic strategies towards modified sugar moieties. (<b>a</b>) Preparation of carbocyclic nicotinamide nucleosides. (<b>b</b>) Prototypic synthesis of an acyclic NAR analog.</p>
Full article ">Figure 8
<p>General strategies for the synthesis of nicotinamide derived C-ribonucleosides. After formation of the nucleosidic bond, an alkylation of the ring nitrogen is required for redox activity.</p>
Full article ">Figure 9
<p>Proposed strategy for synthesis of NCRs by the phosphoramidite approach. The Zincke reaction enables access to partially protected nucleosides from pre-functionalized sugar moieties.</p>
Full article ">Figure 10
<p>Hypothetic outcome of chemical solid phase synthesis using NARH (<b>left</b>) or NAR (<b>right</b>) phosphoramidite. Dashed arrows point to possible degradation products; X: O, CH<sub>2</sub>.</p>
Full article ">
19 pages, 400 KiB  
Review
Perspectives for Photocatalytic Decomposition of Environmental Pollutants on Photoactive Particles of Soil Minerals
by Agnieszka Sosnowska, Kinga I. Hęclik, Joanna B. Kisała, Monika Celuch and Dariusz Pogocki
Materials 2024, 17(16), 3975; https://doi.org/10.3390/ma17163975 - 9 Aug 2024
Viewed by 470
Abstract
The literature shows that both in laboratory and in industrial conditions, the photocatalytic oxidation method copes quite well with degradation of most environmental toxins and pathogenic microorganisms. However, the effective utilization of photocatalytic processes for environmental decontamination and disinfection requires significant technological advancement [...] Read more.
The literature shows that both in laboratory and in industrial conditions, the photocatalytic oxidation method copes quite well with degradation of most environmental toxins and pathogenic microorganisms. However, the effective utilization of photocatalytic processes for environmental decontamination and disinfection requires significant technological advancement in both the area of semiconductor material synthesis and its application. Here, we focused on the presence and “photocatalytic capability” of photocatalysts among soil minerals and their potential contributions to the environmental decontamination in vitro and in vivo. Reactions caused by sunlight on the soil surface are involved in its normal redox activity, taking part also in the soil decontamination. However, their importance for decontamination in vivo cannot be overstated, due to the diversity of soils on the Earth, which is caused by the environmental conditions, such as climate, parent material, relief, vegetation, etc. The sunlight-induced reactions are just a part of complicated soil chemistry processes dependent on a plethora of environmental determinates. The multiplicity of affecting factors, which we tried to sketch from the perspective of chemists and environmental scientists, makes us rather skeptical about the effectiveness of the photocatalytic decontamination in vivo. On the other hand, there is a huge potential of the soils as the alternative and probably cheaper source of useful photocatalytic materials of unique properties. In our opinion, establishing collaboration between experts from different disciplines is the most crucial opportunity, as well as a challenge, for the advancement of photocatalysis. Full article
(This article belongs to the Section Catalytic Materials)
18 pages, 1406 KiB  
Review
Antioxidant Potential of Exosomes in Animal Nutrition
by Hengyu Jin, Jianxin Liu and Diming Wang
Antioxidants 2024, 13(8), 964; https://doi.org/10.3390/antiox13080964 - 8 Aug 2024
Viewed by 300
Abstract
This review delves into the advantages of exosomes as novel antioxidants in animal nutrition and their potential for regulating oxidative stress. Although traditional nutritional approaches promote oxidative stress defense systems in mammalian animals, several issues remain to be solved, such as low bioavailability, [...] Read more.
This review delves into the advantages of exosomes as novel antioxidants in animal nutrition and their potential for regulating oxidative stress. Although traditional nutritional approaches promote oxidative stress defense systems in mammalian animals, several issues remain to be solved, such as low bioavailability, targeted tissue efficiency, and high-dose by-effect. As an important candidate offering regulation opportunities concerned with cellular communication, disease prevention, and physiology regulation in multiple biological systems, the potential of exosomes in mediating redox status in biological systems has not been well described. A previously reported relationship between redox system regulation and circulating exosomes suggested exosomes as a fundamental candidate for both a regulator and biomarker for a redox system. Herein, we review the effects of oxidative stress on exosomes in animals and the potential application of exosomes as antioxidants in animal nutrition. Then, we highlight the advantages of exosomes as redox regulators due to their higher bioavailability and physiological heterogeneity-targeted properties, providing a theoretical foundation and feed industry application. Therefore, exosomes have shown great potential as novel antioxidants in the field of animal nutrition. They can overcome the limitations of traditional antioxidants in terms of dosage and side effects, which will provide unprecedented opportunities in nutritional management and disease prevention, and may become a major breakthrough in the field of animal nutrition. Full article
(This article belongs to the Special Issue Novel Antioxidants for Animal Nutrition—2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The mechanism of exosome production. Cells produce vesicles through endocytosis, which fuse to form early endosomes and gradually become late endosomes, accompanied by the mediation of Golgi apparatus, endoplasmic reticulum, and nucleus. Subsequently, the late endosomes produce many luminal vesicles (ILVs) in the cytoplasm, which gradually evolve into multi-vesicle bodies (MVBs). Finally, it is often accompanied by Endosomal Sorting Complexes Required for Transport (ESCRT) protein complexes, which are released outside the cell to form exosomes.</p>
Full article ">Figure 2
<p>Multiple pathways through which internal components of exosomes alleviate oxidative stress.</p>
Full article ">Figure 3
<p>The potential of milk exosomes in the field of animal nutrition.</p>
Full article ">
15 pages, 5164 KiB  
Article
VO Supported on Functionalized CNTs for Oxidative Conversion of Furfural to Maleic Anhydride
by Pedro Rodríguez, Carolina Parra, J. Noe Díaz de León, Alejandro Karelovic, Sebastian Riffo, Carla Herrera, Gina Pecchi and Catherine Sepúlveda
Catalysts 2024, 14(8), 510; https://doi.org/10.3390/catal14080510 - 7 Aug 2024
Viewed by 384
Abstract
Commercial non-functionalized (CNTs) and functionalized carbon nanotubes (CNT-COOH and CNT-NH2) were used as supports to synthesize vanadium-supported catalysts to be used in the gas phase partial oxidation of furfural towards maleic anhydride (MA). The CNTs and the VO2-V2 [...] Read more.
Commercial non-functionalized (CNTs) and functionalized carbon nanotubes (CNT-COOH and CNT-NH2) were used as supports to synthesize vanadium-supported catalysts to be used in the gas phase partial oxidation of furfural towards maleic anhydride (MA). The CNTs and the VO2-V2O5/CNTs, so-called VO/CNT catalysts, were characterized by AAS, TGA, XRD, N2 adsorption isotherms at −196 °C, Raman, NH3-TPD and XPS. The surface area values, TGA and XRD results indicate that the larger thermal stability and larger dispersion of vanadium species is reached for the VO/CNT-NH2 catalyst. XPS indicates presence of surface VO2 and V2O5 species for the non-functionalized (CNT) and functionalized (CNT-COOH and CNT-NH2) catalysts, with a large interaction of the functional group with the surface vanadium species only for the VO/CNT-NH2 catalyst. The catalytic activity, evaluated in the range 305 °C to 350 °C, indicates that CO, CO2 and MA yield (%) and MA productivity are associated to the redox properties of the vanadium species, the oxygen exchange ability of the support and the vanadium–support interaction. For the reaction temperatures between 320 °C and 335 °C, the maximum MA yield (%) is found in the functionalized VO/CNT-COOH and VO/CNT-NH2 catalysts. This behavior is attributed to a decreased oxidation capability of the CNT with the functionalization. In addition, VO/CNT-NH2 is the more active and selective catalyst for MA productivity at 305 °C and 320 °C, which is related to the greater interaction of the surface vanadium species with the -NH2 group, which enhances the redox properties and stabilization of the VO2 and V2O5 surface active sites. Recycling at 350 °C resulted in 100% furfural conversion for all catalysts and a similar MA yield (%) compared to the fresh catalyst, indicating no loss of surface active sites. Full article
(This article belongs to the Special Issue Catalytic Conversion of Biomass to Chemicals)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Thermal TG (solid line) and DTG (dotted line) profiles for CNTs and VO/CNTs catalysts.</p>
Full article ">Figure 1 Cont.
<p>Thermal TG (solid line) and DTG (dotted line) profiles for CNTs and VO/CNTs catalysts.</p>
Full article ">Figure 2
<p>XRD pattern for CNTs and VO/CNT catalysts.</p>
Full article ">Figure 3
<p>NH<sub>3</sub>-TPD profiles for CNTs and VO/CNT catalysts.</p>
Full article ">Figure 4
<p>XP spectra of O 1s for CNTs and VO/CNTs catalysts (light green C-O; green C=O; light blue O-V<sup>V</sup>; blue O-V<sup>IV</sup>).</p>
Full article ">Figure 5
<p>XP spectra of N 1s for CNTs and VO/CNT catalysts (sky blue N-cationic, light blue N-amine and blue N-amide).</p>
Full article ">Figure 6
<p>XP spectra of V 2p for VO/CNT catalysts (light blue V<sup>V</sup> and blue V<sup>IV</sup>).</p>
Full article ">Figure 7
<p>Yield of MA, CO and CO<sub>2</sub> for VO/CNT catalysts.</p>
Full article ">Figure 8
<p>Productivity of MA for VO/CNT catalysts.</p>
Full article ">
18 pages, 4791 KiB  
Review
Polypyrrole Derivatives: Preparation, Properties and Application
by Lu Hao, Changyi Dong and Demei Yu
Polymers 2024, 16(16), 2233; https://doi.org/10.3390/polym16162233 - 6 Aug 2024
Viewed by 518
Abstract
Polypyrrole (PPy) has attracted widespread attention due to its excellent environmental stability, high conductivity, simple synthesis, good biocompatibility, and reversible redox properties. PPy derivatives not only inherit the advantages of polypyrrole, but also have some unique properties. The side and N-site substitution of [...] Read more.
Polypyrrole (PPy) has attracted widespread attention due to its excellent environmental stability, high conductivity, simple synthesis, good biocompatibility, and reversible redox properties. PPy derivatives not only inherit the advantages of polypyrrole, but also have some unique properties. The side and N-site substitution of PPy can not only yield polymers with good solubility, but it also endows polymers with special functionalities by controlling the introduced functional groups. The performance of copolymers can also be adjusted by the type of monomer or polymerization ratio. In this review, an overview of the different types, main preparation methods, and the application prospects of PPy derivatives reported to date are summarized and presented. The current challenges and future opportunities in this research area are also prospected. Full article
(This article belongs to the Section Polymer Applications)
Show Figures

Figure 1

Figure 1
<p>The synthesis of polypyrrole 3,4-dicarboxylic acid [<a href="#B15-polymers-16-02233" class="html-bibr">15</a>]. Copyright 2019 Polymer.</p>
Full article ">Figure 2
<p>The preparation process of pyrrole derivative monomers by our groups [<a href="#B33-polymers-16-02233" class="html-bibr">33</a>,<a href="#B34-polymers-16-02233" class="html-bibr">34</a>,<a href="#B35-polymers-16-02233" class="html-bibr">35</a>].</p>
Full article ">Figure 3
<p>A fluorescent pyrrole derivative bearing a dansyl substituent [<a href="#B36-polymers-16-02233" class="html-bibr">36</a>]. Copyright 2014 Electrochimica Acta.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic illustration of one-pot electropolymerization using the PEO-b-PMA(Az) template to fabricate laterally integrated conducting polymer nanowires. (<b>b</b>) BF-TEM image of an ultrathin section of PEO114-b-PMA(Az)67 template after one-pot electropolymerization of Py and BiTh. (<b>c</b>) HAADF-STEM image of the ultrathin section [<a href="#B48-polymers-16-02233" class="html-bibr">48</a>]. Copyright 2015 Chemistry of Materials.</p>
Full article ">Figure 5
<p>The effect of N-α substitution on the formation of electroactive PPy films [<a href="#B83-polymers-16-02233" class="html-bibr">83</a>]. Copyright 2014 Journal of Physical Chemistry C.</p>
Full article ">Figure 6
<p>Schematic illustrating the formation of PNVPY nanoparticles [<a href="#B46-polymers-16-02233" class="html-bibr">46</a>]. Copyright 2016 RSC Advances.</p>
Full article ">Figure 7
<p>The mechanism of various components in composite materials [<a href="#B93-polymers-16-02233" class="html-bibr">93</a>]. Copyright 2020 Electrochimica Acta.</p>
Full article ">Figure 8
<p>Antibacterial schematic diagram of PPy-g-CS and COP [<a href="#B102-polymers-16-02233" class="html-bibr">102</a>]. Copyright 2020 Carbohydrate Polymers.</p>
Full article ">Figure 9
<p>Schematic representation of the designed SPE holder and the principle of MIPNs-based Gly sensor [<a href="#B129-polymers-16-02233" class="html-bibr">129</a>]. Copyright 2021 Biosensors and Bioelectronics.</p>
Full article ">Figure 10
<p>(<b>a</b>,<b>b</b>) 3D and (<b>c</b>) SEM images of the micropatterns for (<b>a</b>) PMAEPy and (<b>b</b>,<b>c</b>) OMADDPy film via EWSF [<a href="#B140-polymers-16-02233" class="html-bibr">140</a>]. Copyright 2019 Reactive and Functional Polymers.</p>
Full article ">
17 pages, 4945 KiB  
Article
Metal–Organic Framework-Derived Rare Earth Metal (Ce-N-C)-Based Catalyst for Oxygen Reduction Reactions in Dual-Chamber Microbial Fuel Cells
by Shaik Ashmath, Hao Wu, Shaik Gouse Peera and Tae-Gwan Lee
Catalysts 2024, 14(8), 506; https://doi.org/10.3390/catal14080506 - 5 Aug 2024
Viewed by 425
Abstract
Pt supported on carbon (Pt/C) is deemed as the state-of-the-art catalyst towards oxygen reduction reactions (ORRs) in chemical and biological fuel cells. However, due to the high cost and scarcity of Pt, researchers have focused on the development of Earth-abundant non-precious metal catalysts, [...] Read more.
Pt supported on carbon (Pt/C) is deemed as the state-of-the-art catalyst towards oxygen reduction reactions (ORRs) in chemical and biological fuel cells. However, due to the high cost and scarcity of Pt, researchers have focused on the development of Earth-abundant non-precious metal catalysts, hoping to replace the traditional Pt/C catalyst and successfully commercialize the chemical and biological fuel cells. In this regard, electrocatalysts made of transition metals emerged as excellent candidates for ORRs, especially the electrocatalysts made of Fe and Co in combination with N-doped carbons, which produce potentially active M-N4-C (M=Co, Fe) ORR sites. At present, however, the transition metal-based catalysts are popular; recently, electrocatalysts made of rare earth metals are emerging as efficient catalysts, due to the fact that rare earth metals also have the potential to form rare earth metal-N4-C active sites, just like transition metal Fe-N4-C/Co-N4-C. In addition, mixed valance states and uniqueness of f-orbitals of the rare earth metals are believed to improve the redox properties of the catalyst that helps in enhancing ORR activity. Among the rare earth metals, Ce is the most interesting element that can be explored as an ORR electrocatalyst in combination with the N-doped carbon. Unique f-orbitals of Ce can induce distinctive electronic behavior to the catalyst that helps to form stable coordination structures with N-doped carbons, in addition to its excellent ability to scavenge the OH produced during ORRs, therefore helping in catalyst stability. In this study, we have synthesized Ce/N-C catalysts by a metal–organic framework and pyrolysis strategy. The ORR activity of Ce/N-C catalysts has been optimized by systematically increasing the Ce content and performing RDE studies in 0.1 M HClO4 electrolyte. The Ce/N-C catalyst has been characterized systematically by both physicochemical and electrochemical characterizations. The optimized Ce/N-C-3 catalyst exhibited a half-wave potential of 0.68 V vs. RHE. In addition, the Ce/N-C-3 catalyst also delivered acceptable stability with a loss of 70 mV in its half-wave potential when compared to 110 mV loss for Pt/C (10 wt.%) catalyst, after 5000 potential cycles. When Ce/N-C-3 is used as a cathode catalyst in dual-chamber microbial fuel cells, it delivered a volumetric power density of ~300 mW m−3, along with an organic matter degradation of 74% after continuous operation of DCMFCs for 30 days. Full article
(This article belongs to the Special Issue Recent Advances in Energy-Related Materials in Catalysts, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction images of the N/C and Ce/N-C-3 catalysts. (<b>b</b>–<b>d</b>) SEM images of the Ce/ZIF precursor.</p>
Full article ">Figure 2
<p>(<b>a</b>–<b>d</b>) TEM images of the Ce/N-C-3 catalysts. (<b>e</b>) HAADF image and corresponding elemental mapping of (<b>f</b>) carbon, (<b>g</b>) oxygen, (<b>h</b>) nitrogen, (<b>i</b>) cerium, (<b>j</b>) overall mapping of all the elements in the Ce/N-C-3 catalysts, and (<b>k</b>) line-mapping image and (<b>l</b>) corresponding line mapping of the elements in a particular region of the Ce/N-C-3 catalyst.</p>
Full article ">Figure 3
<p>Deconvoluted spectra of (<b>a</b>) C1s, (<b>b</b>) N 1s, and (<b>c</b>) Ce 3d elements in the CE/N-C-3 catalysts.</p>
Full article ">Figure 4
<p>(<b>a</b>) LSV curves of the Ce/N-C-1,2,3 and four catalysts along with N-C catalyst. (<b>b</b>) LSV comparison of Ce/N-C-3 and Pt/C (10 wt.%) catalyst. (<b>c</b>) Tafel slopes of the Ce/N-C-3 and Pt/C (10 wt.%) catalyst. Electrolyte: O<sub>2</sub>-saturated 0.1 M HClO<sub>4</sub>, scan rate 10 mV sec<sup>−1</sup>, 1600 rpm.</p>
Full article ">Figure 5
<p>LSV curves recorded at different rpms of the catalysts and their corresponding K-L plots for the N-C (<b>a</b>,<b>d</b>), Ce/N-C-3 (<b>b</b>,<b>e</b>), and Pt/C (<b>c</b>,<b>f</b>) catalysts.</p>
Full article ">Figure 6
<p>Stability analysis of the Pt/C and Ce/N-C-3 catalysts. Cyclic voltammograms of the (<b>a</b>) Pt/C and (<b>c</b>) Ce/N-C-3 catalysts. (<b>b</b>) Linear sweep voltammetry curves of the (<b>b</b>) Pt/C and (<b>d</b>) Ce/N-C-3 catalysts. The stability conditions are 0.1 M HClO<sub>4</sub> aqueous solution as the electrolyte. Scan rate for CV cycling test 50 mV sec<sup>−1</sup>; LSV: scan rate 10 mV sec<sup>−1</sup>; 1600 rpm in O<sub>2</sub>-saturated 0.1 M HClO<sub>4.</sub> All the tests were conducted at 25 °C.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic operation of the dual-chamber microbial fuel cells set up and used in this study: (<b>b</b>) OCV values of Pt/C and Ce/N-C-3 catalysts; (<b>c</b>,<b>d</b>) i–v curves for Pt/C and Ce/N-C-3 catalysts; (<b>e</b>) TOC analysis of the anolyte of the microbial fuel cells with Pt/C and Ce/N-C-3 as the cathode catalysts.</p>
Full article ">Scheme 1
<p>Schematic representation of the Ce/ZIF and Ce/N-C- catalyst synthesis and their ORR and MFC performance analysis.</p>
Full article ">
17 pages, 5617 KiB  
Article
Impact of Thermochemical Treatments on Electrical Conductivity of Donor-Doped Strontium Titanate Sr(Ln)TiO3 Ceramics
by Aleksandr Bamburov, Ekaterina Kravchenko and Aleksey A. Yaremchenko
Materials 2024, 17(15), 3876; https://doi.org/10.3390/ma17153876 - 5 Aug 2024
Viewed by 390
Abstract
The remarkable stability, suitable thermomechanical characteristics, and acceptable electrical properties of donor-doped strontium titanates make them attractive materials for fuel electrodes, interconnects, and supports of solid oxide fuel and electrolysis cells (SOFC/SOEC). The present study addresses the impact of processing and thermochemical treatment [...] Read more.
The remarkable stability, suitable thermomechanical characteristics, and acceptable electrical properties of donor-doped strontium titanates make them attractive materials for fuel electrodes, interconnects, and supports of solid oxide fuel and electrolysis cells (SOFC/SOEC). The present study addresses the impact of processing and thermochemical treatment conditions on the electrical conductivity of SrTiO3-derived ceramics with moderate acceptor-type substitution in a strontium sublattice. A-site-deficient Sr0.85La0.10TiO3−δ and cation-stoichiometric Sr0.85Pr0.15TiO3+δ ceramics with varying microstructures and levels of reduction have been prepared and characterized by XRD, SEM, TGA, and electrical conductivity measurements under reducing conditions. The analysis of the collected data suggested that the reduction process of dense donor-doped SrTiO3 ceramics is limited by sluggish oxygen diffusion in the crystal lattice even at temperatures as high as 1300 °C. A higher degree of reduction and higher electrical conductivity can be obtained for porous structures under similar thermochemical treatment conditions. Metallic-like conductivity in dense reduced Sr0.85La0.10TiO3−δ corresponds to the state quenched from the processing temperature and is proportional to the concentration of Ti3+ in the lattice. Due to poor oxygen diffusivity in the bulk, dense Sr0.85La0.10TiO3−δ ceramics remain redox inactive and maintain a high level of conductivity under reducing conditions at temperatures below 1000 °C. While the behavior and properties of dense reduced Sr0.85Pr0.15TiO3+δ ceramics with a large grain size (10–40 µm) were found to be similar, decreasing grain size down to 1–3 µm results in an increasing role of resistive grain boundaries which, regardless of the degree of reduction, determine the semiconducting behavior and lower total electrical conductivity of fine-grained Sr0.85Pr0.15TiO3+δ ceramics. Oxidized porous Sr0.85Pr0.15TiO3+δ ceramics exhibit faster kinetics of reduction compared to the Sr0.85La0.10TiO3−δ counterpart at temperatures below 1000 °C, whereas equilibration kinetics of porous Sr0.85La0.10TiO3−δ structures can be facilitated by reductive pre-treatments at elevated temperatures. Full article
Show Figures

Figure 1

Figure 1
<p>Comparison of literature data on the electrical conductivity of Sr<sub>0.90</sub>La<sub>0.10</sub>TiO<sub>3±δ</sub> ceramics in reducing H<sub>2</sub>-based atmospheres. Sources: Moos [<a href="#B37-materials-17-03876" class="html-bibr">37</a>], Marina [<a href="#B36-materials-17-03876" class="html-bibr">36</a>], Yaremchenko [<a href="#B35-materials-17-03876" class="html-bibr">35</a>], Li [<a href="#B38-materials-17-03876" class="html-bibr">38</a>], Lv [<a href="#B39-materials-17-03876" class="html-bibr">39</a>], Wang [<a href="#B40-materials-17-03876" class="html-bibr">40</a>], Niwa [<a href="#B41-materials-17-03876" class="html-bibr">41</a>], Park [<a href="#B42-materials-17-03876" class="html-bibr">42</a>], and Hashimoto [<a href="#B43-materials-17-03876" class="html-bibr">43</a>]. See <a href="#materials-17-03876-t001" class="html-table">Table 1</a> for the experimental details.</p>
Full article ">Figure 2
<p>Examples of XRD patterns of S85L10 and S85P15 samples. Note that S85L10-1300 indicates the powder as synthesized in air at 1300 °C, and S85P15-1350 corresponds to the ceramic sample sintered in air at 1350 °C. The notations of other samples are listed in <a href="#materials-17-03876-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 3
<p>SEM images of fractured cross-sections of S85L10 ceramics prepared under different conditions.</p>
Full article ">Figure 4
<p>SEM images of fractured cross-sections of S85P15 ceramics prepared under different conditions.</p>
Full article ">Figure 5
<p>Electrical conductivity of S85L10 ceramics as a function of (<b>A</b>) temperature in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and (<b>B</b>) oxygen partial pressure under reducing conditions at 900 °C.</p>
Full article ">Figure 6
<p>(<b>A</b>) Example of thermogravimetric data recorded for powdered S85L10 (Sr<sub>0.85</sub>La<sub>0.10</sub>TiO<sub>3−δ</sub>) ceramics on cooling in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and subsequent heating/cooling cycle in air. (<b>B</b>) Electrical conductivity of dense S85L10 ceramics vs. fraction of Ti<sup>3+</sup> cations in the titanium sublattice.</p>
Full article ">Figure 7
<p>Relaxation of electrical conductivity of (<b>A</b>) oxidized S85L10-1320 ceramics on reduction and (<b>B</b>) reduced S85L10-1320-R-1300 ceramics on redox cycling at 900 °C.</p>
Full article ">Figure 8
<p>Electrical conductivity of S85P15 ceramics as a function of (<b>A</b>) temperature in 10%H<sub>2</sub>-N<sub>2</sub> atmosphere and (<b>B</b>) oxygen partial pressure under reducing conditions at 900 °C.</p>
Full article ">Figure 9
<p>Changes in oxygen nonstoichiometry of S85P15 (Sr<sub>0.85</sub>Pr<sub>0.15</sub>TiO<sub>3+δ</sub>) ceramics on oxidation in air estimated from the thermogravimetric data.</p>
Full article ">Figure 10
<p>Relaxation of electrical conductivity of porous (<b>A</b>) S85P15-1350 and (<b>B</b>) S85P15-H-1350 samples on redox cycling at 850 °C. S85P15-1350 denotes the sample sintered in air at 1350 °C for 10 h.</p>
Full article ">
21 pages, 3151 KiB  
Article
Establishing the Thermodynamic Cards of Dipine Models’ Oxidative Metabolism on 21 Potential Elementary Steps
by Guang-Bin Shen, Shun-Hang Gao, Yan-Wei Jia, Xiao-Qing Zhu and Bao-Chen Qian
Molecules 2024, 29(15), 3706; https://doi.org/10.3390/molecules29153706 - 5 Aug 2024
Viewed by 544
Abstract
Dipines are a type of important antihypertensive drug as L-calcium channel blockers, whose core skeleton is the 1,4-dihydropyridine structure. Since the dihydropyridine ring is a key structural factor for biological activity, the thermodynamics of the aromatization dihydropyridine ring is a significant feature parameter [...] Read more.
Dipines are a type of important antihypertensive drug as L-calcium channel blockers, whose core skeleton is the 1,4-dihydropyridine structure. Since the dihydropyridine ring is a key structural factor for biological activity, the thermodynamics of the aromatization dihydropyridine ring is a significant feature parameter for understanding the mechanism and pathways of dipine metabolism in vivo. Herein, 4-substituted-phenyl-2,6-dimethyl-3,5-diethyl-formate-1,4-dihydropyridines are refined as the structurally closest dipine models to investigate the thermodynamic potential of dipine oxidative metabolism. In this work, the thermodynamic cards of dipine models’ aromatization on 21 potential elementary steps in acetonitrile have been established. Based on the thermodynamic cards, the thermodynamic properties of dipine models and related intermediates acting as electrons, hydrides, hydrogen atoms, protons, and two hydrogen ions (atoms) donors are discussed. Moreover, the thermodynamic cards are applied to evaluate the redox properties, and judge or reveal the possible oxidative mechanism of dipine models. Full article
(This article belongs to the Section Organic Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The chemical structures of dipines and refined dipine models (DH<sub>2</sub>).</p>
Full article ">Figure 2
<p>(<b>a</b>) The oxidative metabolism process of Nifedipine in vivo. (<b>b</b>) Some common oxidoreductases in vivo.</p>
Full article ">Figure 3
<p>Thermodynamic cards of dipine model (DH<sub>2</sub>) aromatization on 21 potential elementary steps during oxidative metabolism.</p>
Full article ">Figure 4
<p>The p<span class="html-italic">K</span><sub>a</sub> of DH<sup>+</sup> and YH<sup>+</sup>, along with the p<span class="html-italic">K</span><sub>a</sub> of common organic acids in acetonitrile.</p>
Full article ">Figure 5
<p>Oxidation potentials (<span class="html-italic">E</span><sub>ox</sub>) of DH<sub>2</sub>, DH<sup>−</sup>, DH<sup>•</sup>, and D<sup>•−</sup>, as well as the reduction potentials (<span class="html-italic">E</span><sub>red</sub>) of common coenzyme models and electron acceptors in acetonitrile (V vs. Fc).</p>
Full article ">Figure 6
<p>Hydricities of DH<sub>2</sub>, and DH<sup>−</sup>, as well as H<sup>−</sup>-affinities of common coenzyme models and hydride acceptors in acetonitrile (kcal/mol).</p>
Full article ">Figure 7
<p>Thermodynamic hydrogen-atom-donating abilities of DH<sub>2</sub>, DH<sup>−</sup>, DH<sub>2</sub><sup>•+</sup>, DH<sup>•</sup>, and DH’<sup>•</sup>, as well as hydrogen atom affinities of common radicals and coenzyme models in acetonitrile (kcal/mol).</p>
Full article ">Figure 8
<p>Thermodynamic proton-donating abilities of DH<sub>2</sub>, DH<sup>+</sup>, DH<sub>2</sub><sup>•+</sup>, and DH’<sup>•</sup>, as well as proton-donating abilities of common acids in acetonitrile (kcal/mol).</p>
Full article ">Figure 9
<p>Thermodynamic abilities of DH<sub>2</sub> and common hydrogen carriers releasing two hydrogen ions (H<sup>−</sup> + H<sup>+</sup>) in the red brackets, releasing two hydrogen atoms (2H<sup>•</sup>) in the purple brackets, and releasing H<sub>2</sub> in the blue brackets in acetonitrile (kcal/mol).</p>
Full article ">Figure 10
<p>Thermodynamic abilities of five possible elementary steps for DH<sub>2</sub> oxidation.</p>
Full article ">Figure 11
<p>Thermodynamic analysis of possible oxidative process for intermediate DH<sub>2</sub><sup>•+</sup>.</p>
Full article ">Figure 12
<p>Thermodynamic card of iAsc accepting two hydrogen ions (atoms) on nine potential elementary steps.</p>
Full article ">Figure 13
<p>Thermodynamic analysis platform of elementary steps for the redox process between 3H<sub>2</sub> and iAsc without any catalyst in acetonitrile.</p>
Full article ">
13 pages, 2534 KiB  
Article
Electrochemical Probing of Human Liver Subcellular S9 Fractions for Drug Metabolite Synthesis
by Daphne Medina, Bhavana Omanakuttan, Ricky Nguyen, Eman Alwarsh and Charuksha Walgama
Metabolites 2024, 14(8), 429; https://doi.org/10.3390/metabo14080429 - 3 Aug 2024
Viewed by 490
Abstract
Human liver subcellular fractions, including liver microsomes (HLM), liver cytosol fractions, and S9 fractions, are extensively utilized in in vitro assays to predict liver metabolism. The S9 fractions are supernatants of human liver homogenates that contain both microsomes and cytosol, which include most [...] Read more.
Human liver subcellular fractions, including liver microsomes (HLM), liver cytosol fractions, and S9 fractions, are extensively utilized in in vitro assays to predict liver metabolism. The S9 fractions are supernatants of human liver homogenates that contain both microsomes and cytosol, which include most cytochrome P450 (CYP) enzymes and soluble phase II enzymes such as glucuronosyltransferases and sulfotransferases. This study reports on the direct electrochemistry and biocatalytic features of redox-active enzymes in S9 fractions for the first time. We investigated the electrochemical properties of S9 films by immobilizing them onto a high-purity graphite (HPG) electrode and performing cyclic voltammetry under anaerobic (Ar-saturated) and aerobic (O2-saturated) conditions. The heterogeneous electron transfer rate between the S9 film and the HPG electrode was found to be 14 ± 3 s−1, with a formal potential of −0.451 V vs. Ag/AgCl reference electrode, which confirmed the electrochemical activation of the FAD/FMN cofactor containing CYP450-reductase (CPR) as the electron receiver from the electrode. The S9 films have also demonstrated catalytic oxygen reduction under aerobic conditions, identical to HLM films attached to similar electrodes. Additionally, we investigated CYP activity in the S9 biofilm for phase I metabolism using diclofenac hydroxylation as a probe reaction and identified metabolic products using liquid chromatography–mass spectrometry (LC-MS). Investigating the feasibility of utilizing liver S9 fractions in such electrochemical assays offers significant advantages for pharmacological and toxicological evaluations of new drugs in development while providing valuable insights for the development of efficient biosensor and bioreactor platforms. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Representative FTIR spectra of HPG/S9 (red) and HPG/HLM (black) on an ATR diamond crystal.</p>
Full article ">Figure 2
<p>Cyclic voltammograms of (<b>A</b>) HPG/HLM and (<b>B</b>) HPG/S9 biofilms at scan rate 0.5 V s<sup>−1</sup> in anaerobic (Ar-purged) phosphate buffer solution, pH 7.0. The solid line represents the data as acquired, while the broken line shows the background-subtracted data, eliminating the non-faradaic charging currents to enhance visualization of the redox peaks. The Y scale on the left corresponds to the background-subtracted data, while the Y scale on the right corresponds to the data as acquired.</p>
Full article ">Figure 3
<p>Trumpet plots displaying the oxidation and reduction peak potentials with logarithm of scan rate for HPG/HLM (black) and HPG/S9 (red) biofilms in anaerobic (Ar-purged) phosphate buffer solution, pH 7.0.</p>
Full article ">Figure 4
<p>Cyclic voltammograms of HPG/HLM (solid black), HPG/S9 (solid red) and HPG/phospholipid (broken black) films at 0.5 V s<sup>−1</sup> in stirred aerobic (O<sub>2</sub>-saturated) phosphate buffer solution, pH 7.0.</p>
Full article ">Figure 5
<p>Extracted-ion chromatograms of the reaction mixture acquired for <span class="html-italic">m</span>/<span class="html-italic">z</span> 296 (broken line) and <span class="html-italic">m</span>/<span class="html-italic">z</span> 312 (solid line) after 1 h of electrolysis of 100.0 μM diclofenac solution in pH 7 phosphate buffer using HPG/S9 electrodes at −0.6 V vs. Ag/AgCl under a constant oxygen supply.</p>
Full article ">Figure 6
<p>MS-MS spectra of the reaction mixture displaying mass peaks and fragmentation patterns for 4-hydroxydiclofenac (<b>A</b>) and diclofenac (<b>B</b>). The inset in (<b>B</b>) shows the lower intensity <span class="html-italic">m</span>/<span class="html-italic">z</span> 296 molecular ion peak of diclofenac. Bulk electrolysis was conducted using HPG/S9 electrodes at −0.6 V vs. Ag/AgCl, under a constant oxygen supply for 1 h in a 100.0 μM diclofenac solution prepared in pH 7 phosphate buffer.</p>
Full article ">Scheme 1
<p>Preparation of various human liver subcellular fractions.</p>
Full article ">Scheme 2
<p>Proposed catalytic cycle of human CYP enzymes involving electrode as electron donor. The CYP catalytic cycle begins with the substrate (RH) entering the enzyme’s active site, displacing a water molecule, and forming an Fe(III)-RH complex. An electron from NADPH (in this case, from the electrode) reduces the iron heme center to Fe(II)-RH. Subsequent oxygen binding and a second electron transfer create an RH-Fe(III)-O<sub>2</sub> complex. Protonation of this complex forms an RH-Fe(III)-OOH peroxo-complex, which then cleaves to produce a reactive [RH-Fe(IV)=O]<sup>•+</sup> species. This oxo-iron species oxidizes the substrate to yield the hydroxylated product (ROH), returning the enzyme to its Fe(III) state and completing the cycle. Hydrogen peroxide can also facilitate this process by forming the RH-Fe(III)-OOH complex and following a similar pathway, known as the peroxide shunt pathway.</p>
Full article ">
35 pages, 12957 KiB  
Article
3-Methyl Thiophene-Modified Boron-Doped Diamond (BDD) Electrodes as Efficient Catalysts for Phenol Detection—A Case Study for the Detection of Gallic Acid in Three Specific Tea Types
by Dhielnawaaz Abrahams and Priscilla G. L. Baker
Foods 2024, 13(15), 2447; https://doi.org/10.3390/foods13152447 - 2 Aug 2024
Viewed by 482
Abstract
Polymer modification has been established as a cost-effective, simple, in situ method for overcoming some of the inherent disadvantages of boron-doped diamond (BDD) electrodes, and its application has been extended to reliable, low-cost environmental monitoring solutions. The present review focuses on modifying BDD [...] Read more.
Polymer modification has been established as a cost-effective, simple, in situ method for overcoming some of the inherent disadvantages of boron-doped diamond (BDD) electrodes, and its application has been extended to reliable, low-cost environmental monitoring solutions. The present review focuses on modifying BDD electrodes with semi-conductive polymers acting as redox mediators. This article reports on the development of a 3-methyl thiophene-modified boron-doped diamond (BDD/P3MT) sensor for the electrochemical determination of total phenolic compounds (TPCs) in tea samples, using gallic acid (GA) as a marker. GA is a significant polyphenol with various biological activities, making its quantification crucial. Thus, a simple, fast, and sensitive GA sensor was fabricated using the electroanalytical square wave voltammetry (SWV) technique. The sensor utilizes a semi-conductive polymer, 3-methyl thiophene, as a redox mediator to enhance BDD’s sensitivity and selectivity. Electrochemical synthesis was used for polymer deposition, allowing for greater purity and avoiding solubility problems. The BDD/P3MT sensor exhibits good electrochemical properties, including rapid charge transfer and a large electrochemical area, enabling GA detection with a limit of detection of 11 mg/L. The sensor’s response was correlated with TPCs measured by the Folin–Ciocalteu method. Square wave voltammetry (SWV) showed a good linear relationship between peak currents and GA concentrations in a wide linear range of 3–71 mg/L under optimal conditions. The BDD/P3MT sensor accurately measured TPCs in green tea, rooibos tea, and black tea samples, with green tea exhibiting the highest TPC levels. The results demonstrate the potential of the modified BDD electrode for the rapid and accurate detection of phenolic compounds in tea, with implications for quality control and antioxidant activity assessments. The prolific publications of the past decade have established BDD electrodes as robust BDD sensors for quantifying polyphenols. Fruits, vegetables, nuts, plant-derived beverages such as tea and wine, traditional Eastern remedies and various herbal nutritional supplements contain phenolic chemicals. The safety concerns of contaminated food intake are significant health concerns worldwide, as there exists a critical nexus between food safety, nutrition, and food security. It has been well established that green tea polyphenol consumption promotes positive health effects. Despite their potential benefits, consuming high amounts of these polyphenols has sparked debate due to concerns over potential negative consequences. Full article
(This article belongs to the Section Food Analytical Methods)
Show Figures

Figure 1

Figure 1
<p>Number of publications in three scientific databases (Google Scholar, Scopus, and Science Direct) issued from 2013 to March 2023 related to keyword searches for “boron-doped diamond electrode, polymer, and phenol detection”.</p>
Full article ">Figure 2
<p>Cyclic voltammogram of the electro-polymerisation of poly-3-methylthiophene from 0.5 M 3-methyl thiophene at BDD in 0.1 M LiCO<sub>4</sub> in acetonitrile at 50 mV/s for 5 scans.</p>
Full article ">Figure 3
<p>Cyclic voltammograms of poly (3-methyl thiophene) at BDD in 0.1 M LiCO<sub>4</sub> in acetonitrile at different scan rates: from inside to outside 10, 20, 30, 40, 50, 60, 70, 80, 90, and 100 mV/s.</p>
Full article ">Figure 4
<p>The electrochemical synthesis pathway of 3-methyl-polythiophene via cyclic voltammetry.</p>
Full article ">Figure 5
<p>The linear variation in the anodic and cathodic peak currents as a function of the square root of scan rates for poly (3-methyl thiophene) in 0.1 M LiCO<sub>4</sub> in acetonitrile.</p>
Full article ">Figure 6
<p>(<b>a</b>) SEM-EDX obtained for the P3MT thin film. (<b>b</b>) SEM micrograph of the polymer film at high magnification (32 µm).</p>
Full article ">Figure 7
<p>AFM image of the unmodified BDD at 32 × 32 μm at 500 mV. (<b>a</b>) Topographic 2D scan, (<b>b</b>) topographic 3D scan, and (<b>c</b>) deflection scan.</p>
Full article ">Figure 8
<p>AFM image of BDD/P3MT at 32 × 32 μm at 500 mV (<b>a</b>) Topographic 2D scan, (<b>b</b>) topographic 3D scan, and (<b>c</b>) deflection scan.</p>
Full article ">Figure 9
<p>Cyclic voltammograms of 0.1 mM gallic acid in phosphate buffer solution (pH 6.6) at (<b>a</b>) the unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at scan rates of 10–100 mV/s.</p>
Full article ">Figure 10
<p>Calibration plot obtained for the scan rate study by cyclic voltammetry of 0.1 mM gallic acid in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at scan rates of 10–100 mV/s.</p>
Full article ">Figure 11
<p>Electrochemical degradation of gallic acid.</p>
Full article ">Figure 12
<p>Concentration study of 0.1 mM gallic acid (0 mg/L to 1.1467 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV.</p>
Full article ">Figure 13
<p>Calibration curves for the concentration study by CV of 0.1 mM gallic acid (0 mg/L to 1 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV.</p>
Full article ">Figure 14
<p>Concentration study of 0.1 mM gallic acid (0 mg/L to 1 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at scan rate of 50 mV/s.</p>
Full article ">Figure 15
<p>Calibration curve of the concentration study by SWV of 0.1 mM gallic acid (0 mg/L to 1 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at a scan rate of 50 mV/s (n = 3).</p>
Full article ">Figure 16
<p>SWV concentration study of 0.1 mM gallic acid (0 to 440 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at a scan rate of 50 mV/s (n = 3).</p>
Full article ">Figure 17
<p>Calibration curve for 0.1 mM gallic acid response measured by SWV (0 mg/L to 124 mg/L) in phosphate buffer solution (pH 6.6) at the (<b>a</b>) unmodified BDD electrode and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at a scan rate of 50 mV/ (n = 3).</p>
Full article ">Figure 18
<p>SWV concentration study of 0.1 mM gallic acid (0 mg/L to 200 mg/L) in phosphate buffer solution (pH 6.6) at (<b>a</b>) BDD and (<b>b</b>) BDD/P3MT in the potential range from −200 mV to 1500 mV at a scan rate of 10 mV/s.</p>
Full article ">Figure 19
<p>Calibration curve for the concentration response of 0.1 mM gallic acid measured via SWV (0 mg/L to 440 mg/L) in phosphate buffer solution (pH 6.6) at (<b>a</b>) BDD and (<b>b</b>) BDD/P3MT in the potential window of −200 mV to 1500 mV at a scan rate of 10 mV/s (n = 3).</p>
Full article ">Figure 20
<p>Examples of screening actual tea samples with a positive gallic acid indication via SWV for (<b>a</b>) rooibos, (<b>b</b>) black, and (<b>c</b>) green tea infusions at BDD/P3MT in the potential range from −200 mV to 1500 mV at a scan rate of 10 mV/s.</p>
Full article ">Figure 21
<p>Calibration curve for gallic acid obtained by the Folin–Ciocalteu reaction method at 765 nm.</p>
Full article ">
26 pages, 9053 KiB  
Review
Constructing Three-Dimensional Architectures to Design Advanced Copper-Based Current Collector Materials for Alkali Metal Batteries: From Nanoscale to Microscale
by Chunyang Kong, Fei Wang, Yong Liu, Zhongxiu Liu, Jing Liu, Kaijia Feng, Yifei Pei, Yize Wu and Guangxin Wang
Molecules 2024, 29(15), 3669; https://doi.org/10.3390/molecules29153669 - 2 Aug 2024
Viewed by 649
Abstract
Alkali metals (Li, Na, and K) are deemed as the ideal anode materials for next-generation high-energy-density batteries because of their high theoretical specific capacity and low redox potentials. However, alkali metal anodes (AMAs) still face some challenges hindering their further applications, including uncontrollable [...] Read more.
Alkali metals (Li, Na, and K) are deemed as the ideal anode materials for next-generation high-energy-density batteries because of their high theoretical specific capacity and low redox potentials. However, alkali metal anodes (AMAs) still face some challenges hindering their further applications, including uncontrollable dendrite growth and unstable solid electrolyte interphase during cycling, resulting in low Coulombic efficiency and inferior cycling performance. In this regard, designing 3D current collectors as hosts for AMAs is one of the most effective ways to address the above-mentioned problems, because their sufficient space could accommodate AMAs’ volume expansion, and their high specific surface area could lower the local current density, leading to the uniform deposition of alkali metals. Herein, we review recent progress on the application of 3D Cu-based current collectors in stable and dendrite-free AMAs. The most widely used modification methods of 3D Cu-based current collectors are summarized. Furthermore, the relationships among methods of modification, structure and composition, and the electrochemical properties of AMAs using Cu-based current collectors, are systematically discussed. Finally, the challenges and prospects for future study and applications of Cu-based current collectors in high-performance alkali metal batteries are proposed. Full article
(This article belongs to the Special Issue Novel Electrode Materials for Rechargeable Batteries, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Synthesis procedure of Cu@Cu<sub>x</sub>O current collector and Li/Cu@Cu<sub>x</sub>O electrode. (<b>b</b>) Side-view scanning electron microscopy (SEM) images of Cu@Cu<sub>x</sub>O. (<b>c</b>) Galvanostatic voltage profiles of Li/Cu@Cu<sub>x</sub>O and Li/PL–Cu symmetric cells. (<b>d</b>) Cycling performance of Li/Cu@Cu<sub>x</sub>O|LFP and Li/PL–Cu|LFP at 1 C. Reprinted with permission [<a href="#B58-molecules-29-03669" class="html-bibr">58</a>]. Copyright 2023, Wiley-VCH. (<b>e</b>) Schematic illustration of the synthetic procedure of the 3D HPC/CF. (<b>f</b>) SEM images of the 3D HPC/CF. (<b>g</b>) Cyclic stability of Li@3D HPC/CF|LFP and Li@CF|LFP full cell at 2 C. Reprinted with permission [<a href="#B60-molecules-29-03669" class="html-bibr">60</a>]. Copyright 2018, American Chemical Society.</p>
Full article ">Figure 2
<p>(<b>a</b>) Progress of synthesizing the 3D Cu–CuSn and 3D Cu–LiSn–Li electrodes. (<b>b</b>) Cross-sectional SEM images with the corresponding EDS elemental mapping of the 3D Cu–CuSn. (<b>c</b>) Cycling stability of the 3D Cu–LiSn–Li||LFP and Li||LFP batteries at 5 C. (<b>d</b>) Rate capabilities of 3D Cu–LiSn–Li||LFP and Li||LFP full cells. Reprinted with permission [<a href="#B69-molecules-29-03669" class="html-bibr">69</a>]. Copyright 2023, American Chemical Society. (<b>e</b>) Schematic diagram of the preparation of 3DHP Cu. (<b>f</b>) SEM image of the 3DHP Cu from the top view. (<b>g</b>) Galvanostatic cycling performance of Li@3DHP Cu electrode. (<b>h</b>) Rate capabilities of Li@Cu||LFP and Li@3DHP Cu||LFP cells. Reprinted with permission [<a href="#B71-molecules-29-03669" class="html-bibr">71</a>]. Copyright 2021, American Chemical Society.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic illustration of the method of synthesizing 3D interconnected porous Cu foam. (<b>b</b>) SEM image of 3D Cu foam. (<b>c</b>) FIB–SEM image of 3D interconnected porous Cu foam. (<b>d</b>) Cycling performance of 3D Cu foam. (<b>e</b>) Li nucleation and plating/stripping behavior on 3D Cu foam electrode. Reprinted with permission [<a href="#B73-molecules-29-03669" class="html-bibr">73</a>]. Copyright 2023, Elsevier. (<b>f</b>) Process of synthesizing CMMC. (<b>g</b>) SEM images of CMMC. (<b>h</b>) SEM images of CMMC. (<b>i</b>) Cycling performances of Li–CMMC and Li–CFC electrodes. (<b>j</b>) Cycling stability of CMMC–LiǀǀLFP battery at 0.2 C. Reprinted with permission [<a href="#B74-molecules-29-03669" class="html-bibr">74</a>]. Copyright 2023, Wiley-VCH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic diagram of method of synthesizing CuM/Ag. (<b>b</b>) Digital photos of CuM/Ag under optical microscopy. (<b>c</b>) Digital photographs of Li@CuM/Ag. (<b>d</b>) Voltage curves of Li plating on Cu mesh and CuM/Ag. (<b>e</b>) Cycling performance of Li, Li@Cu mesh, and Li@CuM/Ag electrodes. (<b>f</b>) Cyclic stability of Li@CuM/Ag||LCO cell at 2 C. Reprinted with permission [<a href="#B84-molecules-29-03669" class="html-bibr">84</a>]. Copyright 2023, Wiley-VCH. (<b>g</b>) Schematic diagram of Ge NWs synthesis on Cu foil. (<b>h</b>) SEM image of Cu-Ge. (<b>i</b>) Electrochemical performance of the Gu–Ge. (<b>j</b>) Cycling stability of Cu–Ge@Li–NMC811 battery at 0.5 C. Reprinted with permission [<a href="#B91-molecules-29-03669" class="html-bibr">91</a>]. Copyright 2023, Wiley-VCH.</p>
Full article ">Figure 5
<p>(<b>a</b>) Process of synthesizing CuO@Cu nanowire arrays. (<b>b</b>) SEM image of the surface view of M–CuO@Cu nanowire arrays. (<b>c</b>) Cycling stability of CuO@Cu–Li anodes in symmetric cells. (<b>d</b>) Cycling performance of LFP||CuO@Cu–Li full-cells at 1 C. Reprinted with permission [<a href="#B92-molecules-29-03669" class="html-bibr">92</a>]. Copyright 2023, Wiley-VCH. (<b>e</b>) The fabrication of roll-pressed Cu@CuO<sub>x</sub> foams (RCOFs). (<b>f</b>) SEM image of RCOFs. (<b>g</b>) Cycling performances of Li–RCOFs electrode. (<b>h</b>) Cycling performances of Li–RCOFs//LFP cell. Reprinted with permission [<a href="#B94-molecules-29-03669" class="html-bibr">94</a>]. Copyright 2020, Elsevier.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic diagram of the fabrication and lithiation process of PDA@3D Cu. (<b>b</b>) Schematic illustration of Li deposition through PDA layer. (<b>c</b>) Cycling performance of Li@PDA@3D Cu electrodes. Reprinted with permission [<a href="#B95-molecules-29-03669" class="html-bibr">95</a>]. Copyright 2020, Elsevier. (<b>d</b>) Process of synthesizing MSEI@Cu. (<b>e</b>) Cycling performances of Li–MSEI@Cu electrode. (<b>f</b>) Cycling stability of Li–MSEI@Cu||LFP cell at 1 C (1C = 170 mAh g<sup>−1</sup>). Reprinted with permission [<a href="#B97-molecules-29-03669" class="html-bibr">97</a>]. Copyright 2024, The Royal Society of Chemistry.</p>
Full article ">Figure 7
<p>(<b>a</b>) Process of synthesizing Sn@LIG@Cu. (<b>b</b>) Voltage curves of Na deposition on different substrates. (<b>c</b>) The cycle performance of Na@Sn@LIG@Cu||NVP at 1 C. Reprinted with permission [<a href="#B105-molecules-29-03669" class="html-bibr">105</a>]. Copyright 2023, Wiley-VCH. (<b>d</b>) Synthesis procedure of Na<sub>2</sub>Se/Cu@Na composite anode. (<b>e</b>) SEM image of CF/Cu<sub>2</sub>Se. (<b>f</b>) The cycling performances of Na||NVP and Na<sub>2</sub>Se/Cu@Na||NVP full batteries at 10 C. (<b>g</b>) The rate capacity of Na<sub>2</sub>Se/Cu@Na||NVP full batteries. Reprinted with permission [<a href="#B106-molecules-29-03669" class="html-bibr">106</a>]. Copyright 2022, Wiley-VCH.</p>
Full article ">Figure 8
<p>(<b>a</b>) Process of synthesizing Pd/Cu foam and K/Pd/Cu foam. (<b>b</b>) Electrochemical performance of the Pd/Cu foam. (<b>c</b>) Cycling stability of K/Pd/Cu||PB cell at −20 °C. Reprinted with permission [<a href="#B109-molecules-29-03669" class="html-bibr">109</a>]. Copyright 2022, Elsevier. (<b>d</b>) Schematic diagram of the synthesis procedure of CuSe@CF. (<b>e</b>) K deposition behavior on KSEC anode, “synchronized” deposition. (<b>f</b>) K deposition behavior on KSC anode, “top-down” depositional. (<b>g</b>) Galvanostatic voltage profiles of KSC and KSEC electrodes. (<b>h</b>) Cycling performance of KSEC–K|PTCDA cell at 2 C. Reprinted with permission [<a href="#B111-molecules-29-03669" class="html-bibr">111</a>]. Copyright 2023, Wiley-VCH.</p>
Full article ">Figure 9
<p>Outlook and prospective research directions relating to advanced 3D Cu-based current collectors for AMBs. Top left: reprinted with permission from [<a href="#B122-molecules-29-03669" class="html-bibr">122</a>] Copyright 2021, American Chemical Society. Top middle: reprinted with permission from [<a href="#B19-molecules-29-03669" class="html-bibr">19</a>] Copyright 2021, Nature Portfolio. The Royal Society of Chemistry. Top right: reprinted with permission from [<a href="#B123-molecules-29-03669" class="html-bibr">123</a>] Copyright 2024, The Royal Society of Chemistry. Bottom left: reprinted with permission from [<a href="#B124-molecules-29-03669" class="html-bibr">124</a>] Copyright 2023, American Chemical Society. Bottom right: reprinted with permission from [<a href="#B92-molecules-29-03669" class="html-bibr">92</a>] Copyright 2023, Wiley-VCH.</p>
Full article ">
11 pages, 2617 KiB  
Article
Synthesis and Properties of 3,8-Diaryl-2H-cyclohepta[b]furan-2-ones
by Taku Shoji, Daichi Ando, Masayuki Iwabuchi, Tetsuo Okujima, Ryuta Sekiguchi and Shunji Ito
Organics 2024, 5(3), 252-262; https://doi.org/10.3390/org5030013 - 1 Aug 2024
Viewed by 278
Abstract
Synthesis of 3,8-diaryl-2H-cyclohepta[b]furan-2-ones was accomplished by the one-pot procedure involving sequential iodation and Suzuki–Miyaura coupling reactions. The optical and structural characteristics of 3,8-diaryl-2H-cyclohepta[b]furan-2-ones prepared were scrutinized using UV/Vis spectroscopy, theoretical calculations, and single-crystal X-ray crystallography. [...] Read more.
Synthesis of 3,8-diaryl-2H-cyclohepta[b]furan-2-ones was accomplished by the one-pot procedure involving sequential iodation and Suzuki–Miyaura coupling reactions. The optical and structural characteristics of 3,8-diaryl-2H-cyclohepta[b]furan-2-ones prepared were scrutinized using UV/Vis spectroscopy, theoretical calculations, and single-crystal X-ray crystallography. The redox properties of the compounds were also evaluated through cyclic voltammetry (CV) and differential pulse voltammetry (DPV). The findings reveal that the introduction of aryl groups at both the 3- and 8-positions significantly influences the electronic properties of the CHFs, resulting in distinct optical and electrochemical characteristics. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of electrophilic substitution and cross-coupling reactions of CHF derivatives.</p>
Full article ">Figure 2
<p>Synthesis of 3,8-diaryl-CHFs <b>4a</b>–<b>4c</b>.</p>
Full article ">Figure 3
<p>ORTEP drawing of (<b>a</b>) <b>4a</b> (CCDC2145584), (<b>b</b>) <b>4b</b> (CCDC2290744), and (<b>c</b>) <b>4c</b> (CCDC2361752) [<a href="#B29-organics-05-00013" class="html-bibr">29</a>].</p>
Full article ">Figure 4
<p>UV/Vis spectra of <b>4a</b> (blue line), <b>4b</b> (red line), and <b>4c</b> (light-green line) in CH<sub>2</sub>Cl<sub>2</sub>.</p>
Full article ">Figure 5
<p>Frontier Kohn–Sham orbitals and their energy levels of <b>4a</b> (left), <b>4b</b> (center), and <b>4c</b> (right) at the B3LYP/6-31G* level.</p>
Full article ">Figure 5 Cont.
<p>Frontier Kohn–Sham orbitals and their energy levels of <b>4a</b> (left), <b>4b</b> (center), and <b>4c</b> (right) at the B3LYP/6-31G* level.</p>
Full article ">
Back to TopTop