[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

Journals

Article Types

Countries / Regions

Search Results (38)

Search Parameters:
Keywords = rotating ring-disk electrode

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
17 pages, 5148 KiB  
Article
Interaction between Vitamins C and E When Scavenging the Superoxide Radical Shown by Hydrodynamic Voltammetry and DFT
by Francesco Caruso, Jens Z. Pedersen, Sandra Incerpi, Stuart Belli, Raiyan Sakib and Miriam Rossi
Biophysica 2024, 4(2), 310-326; https://doi.org/10.3390/biophysica4020022 - 18 Jun 2024
Cited by 1 | Viewed by 728
Abstract
In this study, we examine the cooperative effect between vitamins C and E that mitigates oxidative stress by using experimental and computational methods. We performed superoxide scavenging experiments on each vitamin individually and their combination using rotating ring–disk electrode voltammetry. The results indicate [...] Read more.
In this study, we examine the cooperative effect between vitamins C and E that mitigates oxidative stress by using experimental and computational methods. We performed superoxide scavenging experiments on each vitamin individually and their combination using rotating ring–disk electrode voltammetry. The results indicate that vitamins E and C together produce more effective scavenging of superoxide as evaluated by a steeper slope in the efficiency graph, −7.2 × 104, compared to that of vitamin E alone, −1.8 × 103, or vitamin C alone, −1.3 × 104. Density Functional Theory calculations agree with our experimental results, and we describe a mechanism for the antioxidant action of individual vitamins E and C, plus the synergistic action when both vitamins interact. This process involves the restoration of vitamin E by vitamin C and includes π-π interactions between superoxide and scavengers. The overall result produces an increase in scavenging superoxide radicals when both vitamins act together. Full article
(This article belongs to the Special Issue Molecular Structure and Simulation in Biological System 2.0)
Show Figures

Figure 1

Figure 1
<p>Chemical structures of vitamin E components. (<b>Top</b>) Tocotrienols: α-Tocotrienol: R<sub>1</sub>, R<sub>2</sub>, R<sub>3</sub> = Me; β-Tocotrienol: R<sub>1</sub> = Me, R<sub>2</sub> = H, R<sub>3</sub> = Me; γ-Tocotrienol: R<sub>1</sub> = H, R<sub>2</sub> = Me, R<sub>3</sub> = Me; δ-Tocotrienol: R<sub>1</sub> = H, R<sub>2</sub> = H, R<sub>3</sub> = Me. (<b>Bottom</b>) Tocopherols have the C<sub>12</sub> long chain saturated: α-tocopherol, R<sub>1</sub>, R<sub>2</sub>, R<sub>3</sub> = Me; β-tocopherol, R<sub>1</sub> = Me, R<sub>2</sub> = H, R<sub>3</sub> = Me; γ-tocopherol, R<sub>1</sub> = H, R<sub>2</sub> = Me, R<sub>3</sub> = Me; δ-tocopherol, R<sub>1</sub> = H, R<sub>2</sub> = H, R<sub>3</sub> = Me.</p>
Full article ">Figure 2
<p>Superoxide scavenges H(hydroxyl) in position 6 of vitamin E-model. Minimized structures of reagent (<b>left</b>), transition state (<b>center</b>), and product (<b>bottom right</b>) are shown. This reaction has ΔG of −0.4 Kcal/mol and E(barrier) of 0.8 Kcal/mol. The transferred H atom is equidistant from both moieties in this TS: O(superoxide)—H = 1.219 Å, O6—H = 1.209 Å. O atoms, red; C atoms, black; H atoms, light grey.</p>
Full article ">Figure 3
<p>TS search of scavenging superoxide by vitamin E-model. ΔG = −0.4 kcal/mol; E(barrier) = 0.8 kcal/mol.</p>
Full article ">Figure 4
<p>A proton is placed near the HO<sub>2</sub> moiety of <a href="#biophysica-04-00022-f002" class="html-fig">Figure 2</a> product, having van der Waals separation of 2.60 Å.</p>
Full article ">Figure 5
<p>Minimization of <a href="#biophysica-04-00022-f004" class="html-fig">Figure 4</a> arrangement shows H<sub>2</sub>O<sub>2</sub> formation, separated 1.654 Å from the O(polyphenol) in position 6, which is longer than the initial condition, 1.498 Å, shown in <a href="#biophysica-04-00022-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 6
<p>After elimination of H<sub>2</sub>O<sub>2</sub> from the previous arrangement (<a href="#biophysica-04-00022-f005" class="html-fig">Figure 5</a>), the remaining species, semiquinone vitamin E-model, was minimized, and a superoxide radical was π-π placed above the aromatic ring (van der Waals distance, 3.50 Å). Its geometry minimization shows shortening between both centroids, 2.899 Å, indicating bond formation. Meanwhile, the initial superoxide bond distance of 1.373 Å shortens to 1.302 Å, which indicates some electron density was directed toward the ring.</p>
Full article ">Figure 7
<p>DFT optimization result after a proton was placed near O6 of the previous arrangement, <a href="#biophysica-04-00022-f006" class="html-fig">Figure 6</a>, at O6-proton distance 2.60 Å. It shows O6-H6 bond formation, 0.976 Å; the C6-O6 bond length, 1.378 Å, is a typical single bond and is longer than the partial double bond in the reacting species shown above, 1.295 Å (<a href="#biophysica-04-00022-f006" class="html-fig">Figure 6</a>). This reaction has a small barrier, observed after a TS search, whose resulting structure is shown in <a href="#app1-biophysica-04-00022" class="html-app">Figure S1</a>.</p>
Full article ">Figure 8
<p>Initial state for vitamin C (<b>left</b>) transferring a H atom to the vitamin E-model semiquinone (<b>right</b>), which previously gave up a proton to superoxide; the total charge of this radical is 0.</p>
Full article ">Figure 9
<p>Vitamin C transfers a hydrogen atom to vitamin E-model semiquinone. Left: Minimum of energy after approaching H(vitamin C) to vitamin E-model semiquinone (obtained after DFT minimization of <a href="#biophysica-04-00022-f008" class="html-fig">Figure 8</a>). Right: Product of reaction. Center: Transition state showing H located midway between both O atoms, 1.215 Å from vitamin C and 1.229 Å from vitamin E. This reaction is characterized by ΔG = −2.7 Kcal/mol and E(barrier) = 0.7 Kcal/mol.</p>
Full article ">Figure 10
<p>Vitamin C scavenging superoxide. Initial state, superoxide is placed at van der Waals separation, 2.60 Å, from the acidic proton of vitamin C.</p>
Full article ">Figure 11
<p>Superoxide forms a HO<sub>2</sub> moiety, well separated from the remaining species. DFT calculation performed in n-hexane, an organic environment that can mimic the membrane cell. Bond distances within HO<sub>2</sub>, O-H = 1.027 Å, O-O = 1.398 Å; separation between both units = 1.617 Å; within the vitamin C derivative C-O = 1.277 Å.</p>
Full article ">Figure 12
<p>A 2nd vitamin C molecule, stick style, was placed near the most exposed O(superoxide) of the product arrangement shown in <a href="#biophysica-04-00022-f011" class="html-fig">Figure 11</a>. DFT showed formation of H<sub>2</sub>O<sub>2</sub> well separated from two vitamin C species, 1.680 Å and 1.698 Å. Therefore, H<sub>2</sub>O<sub>2</sub> is a product of vitamin C scavenging superoxide.</p>
Full article ">Figure 13
<p>Interaction between superoxide and the vitamin C radical shown in <a href="#biophysica-04-00022-f012" class="html-fig">Figure 12</a> (right). Superoxide gives its electron to the scavenger, which becomes a monoanionic quinone-like species. Meanwhile, the short O-O bond distance of the incoming radical, 1.296 Å, may be associated with a leaving O<sub>2</sub> molecule, as shown by the separation between both units, 3.713 Å, longer than the van der Waals separation, 2.80 Å.</p>
Full article ">Figure 14
<p>RRDE voltammograms of vitamin E. Each run is associated with a specific color (e.g., red) and corresponds to the red oxidation curve (top, positive current) detected at the ring electrode and the red reduction curve detected at the disk electrode (bottom, negative current). These individual voltammograms show almost no variation; that is, the superoxide destroyed at the ring electrode has a similar value for all runs, suggesting that vitamin E is not a strong antioxidant.</p>
Full article ">Figure 15
<p>Collection efficiency of vitamin E cyclovoltammetry, y = −0.0012x + 18.352, R<sup>2</sup> = 0.8969 considering all points; y = −0.0018x + 18.391, R<sup>2</sup> = 0.8981, excluding the 640 µL data. In experiments at high concentrations of some antioxidants, for instance, quercetin [<a href="#B41-biophysica-04-00022" class="html-bibr">41</a>], a related decreasing pattern in efficiency is seen, and the last data point is not included.</p>
Full article ">Figure 16
<p>Voltammograms for 0.03 M vitamin C. They are much more separated from each other than in <a href="#biophysica-04-00022-f014" class="html-fig">Figure 14</a>, indicating a stronger antioxidant activity with increasing concentration of vitamin C than vitamin E.</p>
Full article ">Figure 17
<p>Voltammograms for vitamin C aliquots in fixed amount of vitamin E, 640 µL.</p>
Full article ">Figure 18
<p>Different slopes of vitamin C and vitamins C plus E. Vitamin C (red line): y = −13,359x + 20.573 R<sup>2</sup> = 0.9947; vitamin E + vitamin C (blue line): y = −71,724x + 17.833 R<sup>2</sup> = 0.99.</p>
Full article ">Scheme 1
<p>Vitamin E scavenges superoxide. (<b>A</b>) Superoxide approaches vitamin E hydroxyl in position 6; (<b>B</b>) H is captured by superoxide, forming [HO<sub>2</sub>]<sup>−</sup> plus vitamin E semiquinone; (<b>C</b>) a proton is captured by [HO<sub>2</sub>]<sup>−</sup>, forming H<sub>2</sub>O<sub>2</sub>; (<b>D</b>) an additional superoxide interacts π-π with the aromatic ring; (<b>E</b>) an additional proton interacts with the semiquinone, forming η-O<sub>2</sub>-vitamin E-model.</p>
Full article ">
14 pages, 7662 KiB  
Article
Heteroatom-Doped Carbon-Based Catalysts Synthesized through a “Cook-Off” Process for Oxygen Reduction Reaction
by Ruiquan Zhang, Qiongyu Liu, Ming Wan, Zhenhua Yao and Maocong Hu
Processes 2024, 12(2), 264; https://doi.org/10.3390/pr12020264 - 25 Jan 2024
Viewed by 1033
Abstract
The development of efficient and low-cost non-metallic catalysts is of great significance for the oxygen reduction reaction (ORR) in fuel cells. Heteroatom-doped carbon-based catalysts are one of the popular candidates, although their preparation method is still under exploration. In this work, single (CS)-, [...] Read more.
The development of efficient and low-cost non-metallic catalysts is of great significance for the oxygen reduction reaction (ORR) in fuel cells. Heteroatom-doped carbon-based catalysts are one of the popular candidates, although their preparation method is still under exploration. In this work, single (CS)-, double (NCS)-, and triple (NBCS)-heteroatom-doped carbon-based catalysts were successfully prepared by a “cook-off” process. The morphology, elemental composition, and bonding structure of the catalysts were investigated by SEM, TEM, Raman spectra, BET, and XPS. ORR catalytic performance measurements suggested an activity trend of CS < NCS < NBCS, and NBCS demonstrated better methanol resistance and slightly higher stability than the commercial Pt/C catalyst, as evaluated with both rotating disk electrode (RDE) and rotating ring-disk electrode (RRDE) systems. The mechanism for the promoted performance was also proposed based on the conductivity of the catalysts. In this paper, the heteroatoms N, B, and S were co-doped into activated carbon using a simple, fast, and efficient preparation method with high electrical conductivity and also increased active sites, showing high electrocatalytic activity and good stability. This work provides a new approach to preparing highly active non-Pt catalysts for oxygen reduction reactions. Full article
Show Figures

Figure 1

Figure 1
<p>Preparation of heteroatom-doped carbon-based catalysts with a “cook-off” process.</p>
Full article ">Figure 2
<p>SEM images of (<b>a</b>) NBCS; (<b>b</b>) NCS; (<b>c</b>) CS.</p>
Full article ">Figure 3
<p>Raman spectra of NBCS, NCS, and CS.</p>
Full article ">Figure 4
<p>(<b>a</b>) N<sub>2</sub> adsorption–desorption isotherms of CS, NCS, NBCS; (<b>b</b>) BJH pore size distribution of CS, NCS, NBCS.</p>
Full article ">Figure 5
<p>(<b>a</b>) XPS of NBCS, NCS, and CS; (<b>b</b>) O 1s of NBCS, NCS, and CS; (<b>c</b>) C 1s of NBCS, NCS, and CS; (<b>d</b>) S 2p of NBCS, NCS, and CS; (<b>e</b>) N 1s of NBCS and NCS; (<b>f</b>) B 1s of NBCS. The solid line in red is the sum of all the peak fittings while the curves in other colors represent elements with different valence state.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) XPS of NBCS, NCS, and CS; (<b>b</b>) O 1s of NBCS, NCS, and CS; (<b>c</b>) C 1s of NBCS, NCS, and CS; (<b>d</b>) S 2p of NBCS, NCS, and CS; (<b>e</b>) N 1s of NBCS and NCS; (<b>f</b>) B 1s of NBCS. The solid line in red is the sum of all the peak fittings while the curves in other colors represent elements with different valence state.</p>
Full article ">Figure 6
<p>(<b>a</b>) LSV curves of NBCS, NCS, CS, and Pt/C; (<b>b</b>) LSV curves of NBCS at different rotational speeds; (<b>c</b>) K-L plots and the calculated number of transferred electrons of NBCS at various potentials; (<b>d</b>) number of transferred electrons of NBCS, NCS, CS, and Pt/C; (<b>e</b>) Tafel plots of NBCS, NCS, CS, and Pt/C; (<b>f</b>) LSV curves recorded on the RRDEs of NBCS, NCS, CS, and Pt/C; (<b>g</b>) yields and number of electrons transferred to H<sub>2</sub>O<sub>2</sub> of NBCS, NCS, CS, and Pt/C; (<b>h</b>) methanol resistance test of NBCS and Pt/C; (<b>i</b>) stability testing of NBCS.</p>
Full article ">Figure 7
<p>Conductivity of NBCS, NCS, and CS at different pressures.</p>
Full article ">
22 pages, 6994 KiB  
Article
Examining the Antioxidant and Superoxide Radical Scavenging Activity of Anise, (Pimpinella anisum L. Seeds), Esculetin, and 4-Methyl-Esculetin Using X-ray Diffraction, Hydrodynamic Voltammetry and DFT Methods
by Miriam Rossi, Francesco Caruso, Natalie Thieke, Stuart Belli, Alana Kim, Elisabetta Damiani, Camilla Morresi and Tiziana Bacchetti
Pharmaceuticals 2024, 17(1), 67; https://doi.org/10.3390/ph17010067 - 31 Dec 2023
Viewed by 1357
Abstract
Pimpinella anisum L., or anise, is a plant that, besides its nutritional value, has been used in traditional medical practices and described in many cultures in the Mediterranean region. A possible reason for anise’s therapeutic value is that it contains coumarins, which are [...] Read more.
Pimpinella anisum L., or anise, is a plant that, besides its nutritional value, has been used in traditional medical practices and described in many cultures in the Mediterranean region. A possible reason for anise’s therapeutic value is that it contains coumarins, which are known to have many biomedical and antioxidant properties. HPLC analysis in our laboratory of the anise extract shows the presence of the coumarin esculetin. We used a hydrodynamic voltammetry rotating ring–disk electrode (RRDE) method to measure the superoxide scavenging abilities of anise seeds and esculetin, which has marked scavenging activity. A related coumarin, 4-methyl-esculetin, also showed strong antioxidant activity as measured by RRDE. Moreover, this study includes the X-ray crystal structure of esculetin and 4-methyl-esculetin, which reveal the H-bond and the stacking intermolecular interactions of the two coumarins. Coordinates of esculetin crystal structure were used to perform a DFT study to arrive at the mechanism of superoxide scavenging. Besides performing a H(hydroxyl) abstraction in esculetin position 6 by superoxide, the scavenging also includes the presence of a second superoxide radical in a π–π approach. Both rings of esculetin were explored for this attack, but only the pyrone ring was effective. As a result, one product of esculetin scavenging is H2O2 formation, while the second superoxide remains π–π trapped within the pyrone ring to form an esculetin-η-O2 complex. Comparison with other coumarins shows that subtle structural differences in the coumarin framework can imply marked differences in scavenging. For instance, when the catechol moiety of esculetin (position 6,7) is shifted to position 7,8 in 4-methyl-7,8-dihydroxy coumarin, that coumarin shows a superoxide dismutase action, which, beside H2O2 formation, includes the formation and elimination of a molecule of O2. This is in contrast with the products formed through esculetin superoxide scavenging, where a second added superoxide remains trapped, and forms an esculetin-η-O2 complex. Full article
Show Figures

Figure 1

Figure 1
<p>Chromatogram of anise extract monitored at 254 nm (black trace) and at 350 nm (red trace) overlayed using commercial esculetin chromatogram monitored at 350 nm (green trace).</p>
Full article ">Figure 2
<p>Absorbance spectrum of anise sample extracted at 12.3 min elution time from <a href="#pharmaceuticals-17-00067-f001" class="html-fig">Figure 1</a> chromatogram (orange trace, right ordinate axis) overlaying the spectrum of commercial esculetin (blue trace, left ordinate axis).</p>
Full article ">Figure 3
<p>X-ray molecular structure of the esculetin (<b>top</b>) and 4-methyl-esculetin (<b>bottom</b>) single molecule found in the asymmetric unit of the crystal structure. Atoms are colored according to standard scheme (red = O, grey = C and white = H) Displacement ellipsoids are drawn at the 50% probability level. Distances among heavy atoms are shown.</p>
Full article ">Figure 4
<p>Important intermolecular hydrogen bonds in the esculetin crystal structure and listed in <a href="#pharmaceuticals-17-00067-t002" class="html-table">Table 2</a> are shown. Yellow hydrogen bond distances are used to indicate similar H-bonding pattern to 4-methyl-esculetin in Figure 6. Distances refer to oxygen–oxygen separation.</p>
Full article ">Figure 5
<p>Stacking distance: 3.211 Å between best planes of two inversion related molecules, (<b>left</b>); offset stacking down <span class="html-italic">b</span>-axis, (<b>right</b>).</p>
Full article ">Figure 6
<p>Important intermolecular hydrogen bonding and short contacts listed in <a href="#pharmaceuticals-17-00067-t003" class="html-table">Table 3</a> in 4-methyl-esculetin along the <span class="html-italic">c</span>-axis are shown. The distances (Å) are the donor–acceptor values. Hydrogen bonds outlined in yellow are similar those in esculetin in <a href="#pharmaceuticals-17-00067-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 7
<p>Unit cell packing diagrams. (<b>Left</b>), offset stacking of two 4-methyl-esculetin molecules in unit cell, down <span class="html-italic">a</span>-axis. (<b>Right</b>), view down <span class="html-italic">b</span>-axis.</p>
Full article ">Figure 8
<p>Voltammogram for anise extract.</p>
Full article ">Figure 9
<p>Collection efficiency of anise extract as a function of aliquot added.</p>
Full article ">Figure 10
<p>RRDE voltammograms of esculetin at increasing concentrations.</p>
Full article ">Figure 11
<p>Collection efficiency of esculetin as a function of concentration, linear slope = −10.7 × 10<sup>4</sup>, only for blank plus two initial aliquots (square spots; diamond spots not included in calculation).</p>
Full article ">Figure 12
<p>Voltammogram of 4-methyl-esculetin showing total concentration (M).</p>
Full article ">Figure 13
<p>Collection efficiency of 4-methyl-esculetin as a function of concentration, linear slope = −14.2 × 10<sup>4</sup>, considering only blank plus three initial aliquots (square spots; diamond spots not included in calculation).</p>
Full article ">Figure 14
<p>The whole set of aliquots for esculetin (red diamonds) and 4-methyl-esculetin (blue circles), show that both coumarins are able to fully eliminate superoxide radicals around the electrodes. However, 4-methyl-esculetin needs less concentration in the electrovoltaic cell for complete elimination of superoxide (efficiency 0).</p>
Full article ">Figure 15
<p>Energy minimum of esculetin, from X-ray atomic coordinates. Selected bond distances are C6-O6, 1.385 Å. C7-O7, 1.372 Å and C2-O2, 1.235 Å.</p>
Full article ">Figure 16
<p>No π–π interactions with superoxide in the pyrone ring, left, nor with the other ring, right, are observed, as upon DFT minimization, the original O-O bond distance of superoxide, 1.373 Å, and the van der Waals separation between superoxide and ring centroids, 3.50 Å, are not modified.</p>
Full article ">Figure 17
<p>Superoxide approaches H6 (van der Waals separation 2.60 Å) to explore H6 capture by esculetin using DFT.</p>
Full article ">Figure 18
<p>DFT result of DFT minimization of <a href="#pharmaceuticals-17-00067-f017" class="html-fig">Figure 17</a> arrangement. No capture of H6 is seen, as O---H (1.519 Å) is longer than the expected bond length, about 1 Å.</p>
Full article ">Figure 19
<p>Result of an approached second superoxide (Figure DFT4) upon DFT minimization: H6 is captured (H6-O bond length = 1.020 Å) and the π–π-added superoxide is accepted by the heterocycle ring, 3.395 Å.</p>
Full article ">Figure 20
<p>A proton approaches the most exposed oxygen in the HO<sub>2</sub> moiety of Figure DFT5, with van der Waals separation of 2.60 Å.</p>
Full article ">Figure 21
<p>After DFT minimization of <a href="#pharmaceuticals-17-00067-f020" class="html-fig">Figure 20</a> arrangement H<sub>2</sub>O<sub>2</sub> forms.</p>
Full article ">Figure 22
<p>Last DFT shows formation of η-O<sub>2</sub>-esculetin. This results from the sequence (1), after H<sub>2</sub>O<sub>2</sub> elimination from Figure DFT5; (2) proton approaching O6 (2.60 Å, O6-H6 formation (0.978 Å). The final product is the esculetin-η-O<sub>2</sub> species.</p>
Full article ">
7 pages, 192 KiB  
Opinion
Measurements of Dioxygen Formation in Catalytic Electrochemical Water Splitting
by Chandan Kumar Tiwari and Yurii V. Geletii
Catalysts 2024, 14(1), 13; https://doi.org/10.3390/catal14010013 - 22 Dec 2023
Viewed by 1314
Abstract
Water oxidation is a multielectron complex reaction that produces molecular oxygen as the final product. The article addresses the lack of confirmation of oxygen product formation in electrochemical oxygen evolution reaction (OER) studies, despite the extensive research conducted on catalysts for water splitting. [...] Read more.
Water oxidation is a multielectron complex reaction that produces molecular oxygen as the final product. The article addresses the lack of confirmation of oxygen product formation in electrochemical oxygen evolution reaction (OER) studies, despite the extensive research conducted on catalysts for water splitting. It critically evaluates the trend observed in many studies that solely rely on electrochemical methods for OER quantification without confirming the oxygen product via complementary analytical techniques. The omission of measuring evolved oxygen gas leaves a crucial gap in the quantification of the OER process and raises concerns about the validity and accuracy of reported results. Analytical techniques, such as gas chromatography, Rotating Ring-Disk Electrode (RRDE), fluorescence oxygen probes, Clark electrode, and volumetry are critically analyzed and described to ensure the reliability and credibility of voltammetry and bulk electrolysis to provide a more accurate assessment of the OER process. Full article
(This article belongs to the Special Issue Electrocatalytic Water Oxidation, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">
14 pages, 3021 KiB  
Article
Azadiradione, a Component of Neem Oil, Behaves as a Superoxide Dismutase Mimic When Scavenging the Superoxide Radical, as Shown Using DFT and Hydrodynamic Voltammetry
by Raiyan Sakib, Francesco Caruso, Stuart Belli and Miriam Rossi
Biomedicines 2023, 11(11), 3091; https://doi.org/10.3390/biomedicines11113091 - 18 Nov 2023
Cited by 2 | Viewed by 1568
Abstract
The neem tree, Azadirachta indica, belongs to the Meliaceae family, and its use in the treatment of medical disorders from ancient times to the present in the traditional medical practices of Asia, Africa and the Middle East is well-documented. Neem oil, extracted [...] Read more.
The neem tree, Azadirachta indica, belongs to the Meliaceae family, and its use in the treatment of medical disorders from ancient times to the present in the traditional medical practices of Asia, Africa and the Middle East is well-documented. Neem oil, extracted from the seeds of the fruit, is widely used, with promising medicinal benefits. Azadiradione, a principal antioxidant component of the seeds of A. indica, is known to reduce oxidative stress and has anti-inflammatory effects. To directly measure the antioxidant ability of neem oil, we used Rotating Ring Disk Electrode (RRDE) hydrodynamic voltammetry to quantify how it can scavenge superoxide radical anions. The results of these experiments show that neem oil is approximately 26 times stronger than other natural products, such as olive oil, propolis and black seed oil, which were previously measured using this method. Next, computational Density Functional Theory (DFT) methods were used to arrive at a mechanism for the scavenging of superoxide radical anions with azadiradione. Our work indicates that azadiradione is an effective antioxidant and, according to our DFT study, its scavenging of the superoxide radical anion occurs through a reaction mechanism in which azadiradione mimics the antioxidant action of superoxide dismutase (SOD). In this mechanism, analogous to the SOD enzymatic reaction, azadiradione is regenerated, along with the production of two products: hydrogen peroxide and molecular oxygen. This antioxidant process provides an explanation for azadiradione’s more general and protective biochemical effects. Full article
(This article belongs to the Section Drug Discovery, Development and Delivery)
Show Figures

Figure 1

Figure 1
<p>Azadiradione molecular structure.</p>
Full article ">Figure 2
<p>RRDE data for neem oil. The bottom part (negative current) shows the formation of superoxide detected at the disk electrode; the top (positive current) shows that the superoxide detected at the ring electrode decreases after adding neem oil aliquots.</p>
Full article ">Figure 3
<p>Collection efficiency of RRDE neem oil. The ring current/disk current (% Efficiency) at each concentration vs. added amount of neem oil shows a decreasing trend.</p>
Full article ">Figure 4
<p>Collection efficiency of neem oil, limited to the first four data points, shows the linear trend y = −2.271x + 20.469, R<sup>2</sup> = 0.9966.</p>
Full article ">Figure 5
<p>Azadiradione was DFT-minimized, and a proton was placed through van der separation forces using the O(carbonyl) associated with the cyclohexene ring, 2.60 Å.</p>
Full article ">Figure 6
<p>Upon DFT minimization of the azadiradione, as shown in <a href="#biomedicines-11-03091-f005" class="html-fig">Figure 5</a>, the O(carbonyl) captures the added proton without any energy barrier, with an O-H bond distance of 0.970 Å, while the associated C-O bond becomes longer, at 1.365 Å, compared to the other C=O carbonyl, 1.227 Å.</p>
Full article ">Figure 7
<p>Near the proton added to azadiradione in <a href="#biomedicines-11-03091-f006" class="html-fig">Figure 6</a>, a superoxide radical (O-O bond distance 1.373 Å) is initially placed through van der Waals forces, 2.60 Å apart (not shown). Upon DFT minimization, the proton becomes linked to superoxide, forming a HO<sub>2</sub> species, which results in its detachment from the remaining azadiradione neutral radical, O---H distance = 1.624 Å.</p>
Full article ">Figure 8
<p>A second proton is initially placed near the more exposed oxygen atom in HO<sub>2</sub>, 2.60 Å (not shown), making the whole arrangement a 1+ charged radical system (from 2 protons plus the reacted negative superoxide). Upon DFT optimization, H<sub>2</sub>O<sub>2</sub> forms and detaches from the organic moiety, 1.572 Å, with the adjacent carbonyl having an only slightly longer CO bond length, 1.250 Å, than the carbonyl at the opposite end of the molecule, 1.220 Å.</p>
Full article ">Figure 9
<p>To the arrangement shown in <a href="#biomedicines-11-03091-f008" class="html-fig">Figure 8</a>, a 2nd superoxide radical was π-π-posed at the van der Waals separation, 3.50 Å from the center of the ring, making the system neutral and non-radical (not shown). After DFT optimization, the superoxide donated its unpaired electron to the ring, forming a molecule of O<sub>2</sub> (O-O bond 1.269 Å, much shorter than the 1.373 Å distance in the superoxide) that then detached from azadiradione, 3.791 Å. Meanwhile, the separation distance of H<sub>2</sub>O<sub>2</sub> from azadiradione increased, 1.685 Å. Thus, after the reaction of azadiradione with two superoxide radicals (<a href="#biomedicines-11-03091-f007" class="html-fig">Figure 7</a> and <a href="#biomedicines-11-03091-f009" class="html-fig">Figure 9</a>) and two protons (<a href="#biomedicines-11-03091-f006" class="html-fig">Figure 6</a> and <a href="#biomedicines-11-03091-f008" class="html-fig">Figure 8</a>), azadiradione is reformed and becomes ready for an additional cycle of superoxide radical scavenging. The reaction products are H<sub>2</sub>O<sub>2</sub> (<a href="#biomedicines-11-03091-f007" class="html-fig">Figure 7</a>) and O<sub>2</sub> (<a href="#biomedicines-11-03091-f009" class="html-fig">Figure 9</a>). <a href="#biomedicines-11-03091-sch001" class="html-scheme">Scheme 1</a> displays the whole process.</p>
Full article ">Figure 10
<p>Molecular structures of tetracyclic terpenoid neem oil compounds closely related to azadiradione, e.g., containing the cyclohexene-carbonyl moiety: azadirone (<b>A</b>), epoxyazadiradione (<b>B</b>), gedunin (<b>C</b>), nimbolide (<b>D</b>) and zafaral (<b>E</b>).</p>
Full article ">Scheme 1
<p>Azadiradione scavenging of superoxide follows the same pattern as the superoxide dismutase enzyme, Reaction (3). No energy barriers were observed for all reactions. In the first step, the initially van-der-Waals-separated proton and O(carbonyl) are established, and upon DFT geometry optimization, the corresponding ΔG<sub>reaction</sub> is −552.3 kcal/mol (top). Next, a green-colored superoxide is placed (van-der-Waals-separated to join the previously added proton), and C=O-H-O<sub>2</sub> is established (ΔG = −1362.5 kcal/mol) (<b>center right</b>). To the center right species, a second (turquoise-colored) proton is placed by the most exposed O atom in the O<sub>2</sub>-H-O-C moiety, and applying DFT minimization, the H<sub>2</sub>O<sub>2</sub> moiety forms and slightly detaches from O(carbonyl), ΔG = −117.3 kcal/mol (<b>center left</b>). Next, a second (pink-colored) superoxide is π-π-attached to the cyclohexene ring (<b>bottom left</b>), and upon DFT minimization, the superoxide transfers its unpaired electron to the ring, and O<sub>2</sub> is eliminated along with the previously formed H<sub>2</sub>O<sub>2</sub>, ΔG = −17.3 kcal/mol. Thus, the final products are indicated, including reformed azadiradione, ready to start another catalytic cycle (<b>bottom right</b>).</p>
Full article ">
17 pages, 5042 KiB  
Article
Maytenus octogona Superoxide Scavenging and Anti-Inflammatory Caspase-1 Inhibition Study Using Cyclic Voltammetry and Computational Docking Techniques
by Francesco Caruso, Miriam Rossi, Eric Eberhardt, Molly Berinato, Raiyan Sakib, Felipe Surco-Laos and Haydee Chavez
Int. J. Mol. Sci. 2023, 24(13), 10750; https://doi.org/10.3390/ijms241310750 - 28 Jun 2023
Viewed by 1069
Abstract
The relationship between oxidative stress and inflammation is well known, and exogenous antioxidants, primarily phytochemical natural products, may assist the body’s endogenous defense systems in preventing diseases due to excessive inflammation. In this study, we evaluated the antioxidant properties of ethnomedicines from Peru [...] Read more.
The relationship between oxidative stress and inflammation is well known, and exogenous antioxidants, primarily phytochemical natural products, may assist the body’s endogenous defense systems in preventing diseases due to excessive inflammation. In this study, we evaluated the antioxidant properties of ethnomedicines from Peru that exhibit anti-inflammatory activity by measuring the superoxide scavenging activity of ethanol extracts of Maytenus octogona aerial parts using hydrodynamic voltammetry at a rotating ring-disk electrode (RRDE). The chemical compositions of these extracts are known and the interactions of three methide-quinone compounds found in Maytenus octogona with caspase-1 were analyzed using computational docking studies. Caspase-1 is a critical enzyme triggered during the activation of the inflammasome and its actions are associated with excessive release of cytokines. The most important amino acid involved in active site caspase-1 inhibition is Arg341 and, through docking calculations, we see that this amino acid is stabilized by interactions with the three potential methide-quinone Maytenus octogona inhibitors, hydroxytingenone, tingenone, and pristimerin. These findings were also confirmed after more rigorous molecular dynamics calculations. It is worth noting that, in these three compounds, the methide-quinone carbonyl oxygen is the preferred hydrogen bond acceptor site, although tingenone’s other carbonyl group also shows a similar binding energy preference. The results of these calculations and cyclovoltammetry studies support the effectiveness and use of anti-inflammatory ethnopharmacological ethanol extract of Maytenus octogona (L’Héritier) DC. Full article
Show Figures

Figure 1

Figure 1
<p>Two-dimensional canonical structural display of compounds analyzed in this study. Synonyms: hydroxytingenone = beta-hydroxytingenone; maytenin = tingenone.</p>
Full article ">Figure 2
<p>Pose 4 of hydroxytingenone docking, two-dimensional display. The methide-quinone carbonyl is engaged in hydrogen bond with Arg341 and His237, in contrast with the other carbonyl that shows no interactions.</p>
Full article ">Figure 3
<p>More detailed features of pose 4, hydroxytingenone, docking. The donor guanidino group of Arg341 (turquoise colored C atoms) has a double H-bond interaction with O(methide-quinone) carbonyl, 2.636 Å and 2.979 Å. The O(hydroxyl) associated with the methide-quinone inhibitor has a double H-bond interaction, one with H(Cys285), 1.837 Å, and one with H(His237), 2.772 Å (lower left of figure).</p>
Full article ">Figure 4
<p>After molecular dynamics pose 4, hydroxytingenone has calculated binding energy of −13.7 kcal/mol. The H-bond interaction between the donor guanidino group of Arg341 and O(methide-quinone) carbonyl is confirmed and strengthened, 2.109 Å, whereas an H-bond of 1.663 Å by acceptor O(Ser339) replaces the former interactions by Cys285 and His237. A comparison with interactions shown by the inhibitor VX-765 at the crystal structure active site [<a href="#B22-ijms-24-10750" class="html-bibr">22</a>], shows involvement of both amino acids, Arg341 and O(Ser339), also forming hydrogen bonds; the latter interacts with a guanidino HN moiety. Indeed, the role of Arg341 appears very important as it has a double hydrogen bond interaction with two different O(carbonyl) moieties of VX-765.</p>
Full article ">Figure 5
<p>Pose 1. The tingenone Docking 2D, Arg341 forms a hydrogen bond to the non-methide-quinone carbonyl, which also has H-bond interaction with H(His237). The methide-quinone moiety appears free of interactions.</p>
Full article ">Figure 6
<p>Pose 1. The tingenone dynamics of pose 1 has a calculated binding energy of −11.8 kcal/mol. The guanidino group of Arg383 (turquoise colored) establishes a cation-π interaction with the methide-quinone ring, 2.752 Å, whereas the originally docked Arg341 (violet colored) has a double non-chelating H-bond from the guanidino group to the opposite carbonyl, 3.262 Å and 3.088 Å. Interestingly, the cation-π interaction reminds us of the interaction that the same amino acid, Arg383, establishes with an aromatic ring of original inhibitor VX-765 in the crystal structure.</p>
Full article ">Figure 7
<p>Pose 5. Tingenone docking 2D interactions. Arg341 establishes an H-bond to the non-methide-quinone moiety.</p>
Full article ">Figure 8
<p>Pose 5. Tingenone dynamics has a calculated binding energy of −13.0 kcal/mol. It shows a very short H-bond distance of 1.971 Å, from guanidino H(Arg341) to the O(carbonyl) opposed to the methide-quinone carbonyl.</p>
Full article ">Figure 9
<p>Pose 4. Pristimerin docking results, two-dimensional display. Cys285 shows π-sulfur interaction with the methide-quinone ring and Arg341 has a H-bond to the methide-quinone O(carbonyl).</p>
Full article ">Figure 10
<p>Pristimerin pose 4 dynamics after calculating the binding energy, −9.4 kcal. O(methide-quinone-carbonyl) establishes an H-bond to brown colored H(Arg341), 2.370 Å, and to turquoise colored H(Trp340), 2.297 Å, whereas O(Ser339) is 1.524 Å from pristimerin H(hydroxy). Interestingly, the above mentioned O(Ser339) interaction seen in hydroxytingenone, and found in VX-765 crystal structure through an HN moiety, is identical to that seen in pristimerin, supporting a potential role of this ligand and hydroxytingenone at the active site. In addition, the interaction between Trp340, indicated as π-alkyl in docking (<a href="#ijms-24-10750-f009" class="html-fig">Figure 9</a>), and directed to the ring adjacent to the methide-quinone, transforms to hydrogen bonding with O(carbonyl) after dynamics, 2.297 Å.</p>
Full article ">Figure 11
<p>MORUnica (stem extract) RRDE voltammograms. Each run is associated with a specific color, for instance the 70 μL aliquot red line can be observed in the upper part, oxidation curve, detected at the ring electrode, which is consistent with the red line detected at the disk electrode for reduction during the same experiment and located at the bottom of the figure.</p>
Full article ">Figure 12
<p>MORUnica collection efficiency shows a linear behavior, y = −0.008x + 12.807, R<sup>2</sup> = 0.9757, whose slope is associated with the superoxide scavenging capability of the stem extract.</p>
Full article ">Figure 13
<p>MOHUnica (leaf extract) RRDE voltammograms.</p>
Full article ">Figure 14
<p>MOHUnica collection efficiency of the ethanol leaf extract, y = −0.0232x + 10.585, R<sup>2</sup> = 0.994. Compared with the stem extract (<a href="#ijms-24-10750-f012" class="html-fig">Figure 12</a>), the steeper slope in this figure indicates a stronger capability of scavenging superoxide.</p>
Full article ">
17 pages, 6396 KiB  
Article
Antioxidant Properties of Thymoquinone, Thymohydroquinone and Black Cumin (Nigella sativa L.) Seed Oil: Scavenging of Superoxide Radical Studied Using Cyclic Voltammetry, DFT and Single Crystal X-ray Diffraction
by Raiyan Sakib, Francesco Caruso, Sandjida Aktar, Stuart Belli, Sarjit Kaur, Melissa Hernandez and Miriam Rossi
Antioxidants 2023, 12(3), 607; https://doi.org/10.3390/antiox12030607 - 1 Mar 2023
Cited by 10 | Viewed by 3716
Abstract
Black cumin seeds and seed oil have long been used in traditional foods and medicine in South Asian, Middle Eastern and Mediterranean countries and are valuable flavor ingredients. An important ingredient of black cumin is the small molecule thymoquinone (TQ), which manifests low [...] Read more.
Black cumin seeds and seed oil have long been used in traditional foods and medicine in South Asian, Middle Eastern and Mediterranean countries and are valuable flavor ingredients. An important ingredient of black cumin is the small molecule thymoquinone (TQ), which manifests low toxicity and potential therapeutic activity against a wide number of diseases including diabetes, cancer and neurodegenerative disorders. In this study, the antioxidant activities of black seed oil, TQ and a related molecule found in black cumin, thymohydroquinone (THQ), were measured using a direct electrochemical method to experimentally evaluate their superoxide scavenging action. TQ and the black seed oil showed good superoxide scavenging ability, while THQ did not. Density Functional Theory (DFT) computational methods were applied to arrive at a chemical mechanism describing these results, and confirmed the experimental Rotating Ring Disk Electrode (RRDE) findings that superoxide oxidation to O2 by TQ is feasible, in contrast with THQ, which does not scavenge superoxide. Additionally, a thorough inquiry into the unusual cyclic voltammetry pattern exhibited by TQ was studied and was associated with formation of a 1:1 TQ-superoxide radical species, [TQ-O2]•. DFT calculations reveal this radical species to be involved in the π-π mechanism describing TQ reactivity with superoxide. The crystal structures of TQ and THQ were analyzed, and the experimental data reveal the presence of stacking intermolecular interactions that can be associated with formation of the radical species, [TQ-O2]•. All three of these methods were essential for us to arrive at a chemical mechanism that explains TQ antioxidant activity, that incorporates intermolecular features found in the crystal structure and which correlates with the measured superoxide scavenging activity. Full article
(This article belongs to the Special Issue Antioxidants and Oxidative Stability in Fats and Oils)
Show Figures

Figure 1

Figure 1
<p>Asymmetric unit of THQ, (<b>left</b>), and TQ, (<b>right</b>), with atom labeling.</p>
Full article ">Figure 2
<p>TQ molecules have planar quinone rings which have close contacts with other TQ molecules in the same plane of about 3.4 Å. Stacking between these planes is also about 3.4 Å. Hydrogen atoms removed for clarity.</p>
Full article ">Figure 3
<p>Hydrogen bonding among the THQ molecules in the crystal. Hydrogen atoms are not shown for clarity.</p>
Full article ">Figure 4
<p>TQ superoxide scavenging DFT calculations. H atoms are light blue to avoid white background misperception. (<b>A</b>) Minimized TQ molecule obtained after input of corresponding X-ray coordinates. (<b>B</b>) Superoxide was π-π placed (3.50 Å) above ring shown in A, and upon DFT minimization both moieties became more distant, 3.579 Å, while ring C-C bond lengths were modified due to electron transferred from superoxide to the ring and O-O bond length shortens to 1.283 Å. (<b>C</b>) Structure shown in A has a proton van der Waals separated (2.60 Å) from one O(carbonyl). (<b>D</b>) DFT outcome of structure shown in C, plus an additional proton 2.60 Å separated from the second O(carbonyl). (<b>E</b>) DFT outcome of structure shown in D whose charge is 2+, showing two C-OH moieties. (<b>F</b>) A superoxide radical was π-π placed in E, and DFT shows formation of the radical [thymohydroquinone-η-O<sub>2</sub>]•<sup>+</sup>, with separation between centroids of 2.697 Å.</p>
Full article ">Figure 5
<p>An additional superoxide (large stick style) was π-π placed below the structure shown in <a href="#antioxidants-12-00607-f004" class="html-fig">Figure 4</a>F and its DFT minimization shows the neutral final complex [thymohydroquinone-η-O<sub>2</sub>], obtained for process (b).</p>
Full article ">Figure 6
<p>THQ DFT calculations. (<b>A</b>) Geometry minimization of thymohydroquinone X-ray coordinates. (<b>B</b>) Minimization obtained after placing a van der Waals separated superoxide from one H(hydroxyl). (<b>C</b>) Minimization obtained after placing a van der Waals separated HO<sub>2</sub><sup>−</sup> from thymohydroquinone radical. (<b>D</b>) Minimization obtained after placing a π-π van der Waals separated superoxide from THQ, 3.50 Å. Initial superoxide bond length of 1.373 Å had only a slight variation, 1.368 Å. This indicates rejection between both reagents.</p>
Full article ">Figure 7
<p>TQ RRDE cyclovoltammetry.</p>
Full article ">Figure 8
<p>TQ efficiency.</p>
Full article ">Figure 9
<p>TQ collection efficiency for blank plus first four aliquots, whose line equation is y = −21104x + 20.9 (R<sup>2</sup> = 0.9965).</p>
Full article ">Figure 10
<p>CV of TQ 0.00074 M solution (1280 µL aliquot) saturated with the O<sub>2</sub>/N<sub>2</sub> mixture. The star and arrows are described in the text.</p>
Full article ">Figure 11
<p>CV of TQ 0.00074 M solution (1280 µL aliquot) purged with N<sub>2</sub> gas showing evidence of incomplete purging of the O<sub>2</sub>.</p>
Full article ">Figure 12
<p>A cyclovoltammogram of a blank experiment using a classical one working electrode, is taken from [<a href="#B24-antioxidants-12-00607" class="html-bibr">24</a>], where meaning of red lines are described.</p>
Full article ">Figure 13
<p>DFT minimum energy reaction product, obtained after geometry minimization of π-π approach of a van der Waals separated, 3.50 Å, molecular oxygen and the TQ radical [TQ]•<sup>−</sup>. This radical product had a shortened π-π interaction of 2.579 Å and is associated with <span class="html-fig-inline" id="antioxidants-12-00607-i001"><img alt="Antioxidants 12 00607 i001" src="/antioxidants/antioxidants-12-00607/article_deploy/html/images/antioxidants-12-00607-i001.png"/></span> peak oxidation in <a href="#antioxidants-12-00607-f009" class="html-fig">Figure 9</a>, to give TQ plus O<sub>2</sub>.</p>
Full article ">Figure 14
<p>THQ is not able to scavenge superoxide, as no variation is found with increasing concentration of THQ.</p>
Full article ">Figure 15
<p>Black seed oil RRDE electrochemistry.</p>
Full article ">Figure 16
<p>Black seed oil collection efficiency.</p>
Full article ">Figure 17
<p>Black seed oil collection efficiency of blank plus first four aliquots, y = −0.0781x + 20.08, R<sup>2</sup> = 0.9763.</p>
Full article ">
19 pages, 4822 KiB  
Article
Scavenging of Superoxide in Aprotic Solvents of Four Isoflavones That Mimic Superoxide Dismutase
by Sandra Yu, Francesco Caruso, Stuart Belli and Miriam Rossi
Int. J. Mol. Sci. 2023, 24(4), 3815; https://doi.org/10.3390/ijms24043815 - 14 Feb 2023
Cited by 9 | Viewed by 1651
Abstract
Isoflavones are plant-derived natural products commonly found in legumes that show a large spectrum of biomedical activities. A common antidiabetic remedy in traditional Chinese medicine, Astragalus trimestris L. contains the isoflavone formononetin (FMNT). Literature reports show that FMNT can increase insulin sensitivity and [...] Read more.
Isoflavones are plant-derived natural products commonly found in legumes that show a large spectrum of biomedical activities. A common antidiabetic remedy in traditional Chinese medicine, Astragalus trimestris L. contains the isoflavone formononetin (FMNT). Literature reports show that FMNT can increase insulin sensitivity and potentially target the peroxisome proliferator-activated receptor gamma, PPARγ, as a partial agonist. PPARγ is highly relevant for diabetes control and plays a major role in Type 2 diabetes mellitus development. In this study, we evaluate the biological role of FMNT, and three related isoflavones, genistein, daidzein and biochanin A, using several computational and experimental procedures. Our results reveal the FMNT X-ray crystal structure has strong intermolecular hydrogen bonding and stacking interactions which are useful for antioxidant action. Cyclovoltammetry rotating ring disk electrode (RRDE) measurements show that all four isoflavones behave in a similar manner when scavenging the superoxide radical. DFT calculations conclude that antioxidant activity is based on the familiar superoxide σ-scavenging mode involving hydrogen capture of ring-A H7(hydroxyl) as well as the π–π (polyphenol–superoxide) scavenging activity. These results suggest the possibility of their mimicking superoxide dismutase (SOD) action and help explain the ability of natural polyphenols to assist in lowering superoxide concentrations. The SOD metalloenzymes all dismutate O2•− to H2O2 plus O2 through metal ion redox chemistry whereas these polyphenolic compounds do so through suitable hydrogen bonding and stacking intermolecular interactions. Additionally, docking calculations suggest FMNT can be a partial agonist of the PPARγ domain. Overall, our work confirms the efficacy in combining multidisciplinary approaches to provide insight into the mechanism of action of small molecule polyphenol antioxidants. Our findings promote the further exploration of other natural products, including those known to be effective in traditional Chinese medicine for potential drug design in diabetes research. Full article
(This article belongs to the Special Issue 25th Anniversary of IJMS: Advances in Biochemistry)
Show Figures

Figure 1

Figure 1
<p>Single FMNT molecule with atom labeling scheme and showing the −44.3° torsion angle of the twisted (B) phenyl ring.</p>
Full article ">Figure 2
<p>(<b>A</b>) FMNT strong hydrogen bond chain throughout the crystal structure. (<b>B</b>) Intermolecular interactions of offset stacking among FMNT molecules.</p>
Full article ">Figure 3
<p>Molecules are truncated to allow larger views, only (<b>A</b>) is complete. (<b>A</b>) DFT minimized molecule of FMNT after the input of X-ray coordinates, a slight variation of the torsion angle, −34.8°, is seen when compared with the X-ray structure, −44.7°. H atoms belonging to ring A are blue. (<b>B</b>) A superoxide was posed 2.60 Å near FMNT H7 and upon DFT minimization H7, it was captured by superoxide, H7-O(superoxide) = 1.103 Å following H atom transfer (HAT). (<b>C</b>) A proton was van der Waals posed near the most exposed O(superoxide) and captured, 0.983 Å. However, O7-H7 is reformed. 1.044 Å. (<b>D</b>) A second superoxide is π–π posed over ring A, and upon DFT this becomes bound to the ring, 3.038 Å, while H<sub>2</sub>O<sub>2</sub> is formed and well separated, 1.636 Å, from the remaining polyphenol. Calculation in DMSO solvent shows a slightly longer separation between centroids, 3.043 Å. The difference of energy between the reagent and product in the gas phase is −41.9 Kcal/mol. (<b>E</b>) After H<sub>2</sub>O<sub>2</sub> is eliminated, DFT shows still the second superoxide trapped by ring A, with a distance of 3.055 Å between both centroids, shorter than the 3.50 van der Waals separation. (<b>F</b>) A proton is posed by O7, and upon DFT minimization O7-H7 is formed, 0.975 Å. This is a neutral molecule that still comprises the π–π bound second superoxide.</p>
Full article ">Figure 4
<p>After approaching an additional superoxide of σ type close to H7, the π–π bound original superoxide, separated 3.015 Å from ring A (<a href="#ijms-24-03815-f003" class="html-fig">Figure 3</a>F), is expelled from the polyphenol, 5.729 Å, as shown by this unfinished geometry optimization, and FMNT is reformed. It will be shown later that in the equivalent situation genistein behaves differently regarding ring C.</p>
Full article ">Figure 5
<p>Some genistein H atoms are light blue to avoid white background misperception. (<b>A</b>) After posing superoxide 2.60 Å from H7(Hydroxyl), DFT minimization shows hydroxyl elongation, 1.445 Å, and H capture, O(superoxide)-H7 = 1.084 Å (following an H atom transfer (HAT)), forming HO<sub>2</sub><sup>−</sup>. (<b>B</b>) HO<sub>2</sub><sup>−</sup> was posed near a proton and DFT minimization showed H<sub>2</sub>O<sub>2</sub> formation, separated 1.621 Å from O7, not shown, then a second superoxide was π–π posed on top of the A ring, with a separation of 3.50 Å between their centroids. Upon DFT minimization, the centroid separation shortened slightly by 3.477 Å, the superoxide O-O bond length became similar to that in the molecule of O<sub>2</sub>, 1.276 Å, i.e., suggesting transfer of the superoxide electron to the polyphenol. Meanwhile, O7-H7 became closer, 1.573 Å. (<b>C</b>) After the elimination of H<sub>2</sub>O<sub>2</sub> in B, a proton was posed near O7, 2.60 Å, and minimization resulted in restoring the genistein O7-H7 bond. This also strengthened the penetration of O<sub>2</sub> in the ring environment, 2.931 Å. (<b>D</b>) After an additional superoxide was σ type posed to H7, DFT shows the π–π bound molecule of O<sub>2</sub> leaving ring A, 4.398 Å, and relocating above ring C, i.e., having both centroids separated by 3.154 Å. This structure differs from the equivalent one shown in <a href="#ijms-24-03815-f004" class="html-fig">Figure 4</a>, where the molecule of O<sub>2</sub> was completely displaced from FMNT.</p>
Full article ">Figure 6
<p>(<b>A</b>) The π–π interaction between superoxide and daidzein ring A results in the same arrangement as that seen performing a σ interaction, i.e., the original separation between superoxide and ring A centroids, 3.50 Å, becomes 4.997 Å, while H7 is captured by superoxide (following an H atom transfer (HAT)), O(superoxide-H7) = 1.047 Å (1.069 Å in DMSO solvent), and HO<sub>2</sub><sup>−</sup> forms, well separated from O7, 1.512 Å (1.445 Å in DMSO). The difference of energy between the reagent and product in the gas phase is −181.7 Kcal/mol. (<b>B</b>) After posing the more exposed oxygen atom of HO<sub>2</sub><sup>−</sup> near a proton, DFT minimization shows H<sub>2</sub>O<sub>2</sub> formation, well separated from O7, 1.619 Å. (<b>C</b>) After elimination of H<sub>2</sub>O<sub>2</sub> and posing a π–π superoxide over ring A, DFT minimization shows the formation of O<sub>2</sub>, seen by an O-O bond length of 1.264 Å, separated by 3.494 Å from ring A centroid, i.e., the superoxide electron has been transferred to the polyphenol system. Therefore, after posing a proton 2.60 Å near O7 and applying DFT minimization, a molecule of O<sub>2</sub> results in being eliminated and daidzdein becomes reformed ready for performing another SOD cycle, not shown.</p>
Full article ">Figure 7
<p>Biochanin A is reformed after acting like SOD, and includes a π–π bound molecule of O<sub>2</sub>. This O<sub>2</sub>-η-biochanin complex is able to initiate another cycle of scavenging superoxide. This structure is closely related to that shown in <a href="#ijms-24-03815-f003" class="html-fig">Figure 3</a>F.</p>
Full article ">Figure 8
<p>FMNT scavenges superoxide radicals in DMSO solution.</p>
Full article ">Figure 9
<p>Genistein RRDE scavenging efficiency. The first two aliquots were not effective in scavenging superoxide and were omitted for line calculation.</p>
Full article ">Figure 10
<p>Daidzein RRDE cyclic voltammetry collection efficiency.</p>
Full article ">Figure 11
<p>Biochanin collection efficiency.</p>
Full article ">Figure 12
<p>Standard dynamic cascade shows Cys285 forming a strong π-alkyl bond with FMNT’s pyrone and A rings. FMNT forms a sort of basket arrangement at the active site.</p>
Full article ">Figure 13
<p>Dynamics calculation of FMNT shows its stabilization in the ligand-binding pocket by a combination of van der Waals interactions and π-alkyl bonds, including the important amino acids Cys285, Arg288 and Ile341.</p>
Full article ">Scheme 1
<p>Formononetin (FMNT) interactions with superoxide. Bottom (σ attack): H(hydroxyl) in position 7 is captured (green arrow). Top (π attack towards the three ring centers): only ring A is effective (turquoise arrow), and the superoxide radical is directed towards H7, i.e., as when originated from σ scavenging. Interactions with rings B and C result in superoxide rejection.</p>
Full article ">Scheme 2
<p>FMNT scavenging of superoxide. No barriers were present in these DFT geometry minimizations.</p>
Full article ">Scheme 3
<p>Genistein interactions with superoxide. Bottom (σ attack): H(hydroxyl) in position 7 is captured (green arrow), in contrast with H(hydroxyl) in position 4′ which establishes a H-bond, and with H5, resulting in superoxide rejection. Top (π–π attack with the three ring centers): ring A is positively concerned for scavenging and the superoxide radical is directed towards H7, i.e., as when originated from σ-scavenging. Interactions with ring B result in H-bond to H4′, i.e., behaving as in the related H4′ σ-scavenging. With ring C (pyrone ring), a genistein-η-O<sub>2</sub> complex is established, turquoise arrow (*).</p>
Full article ">Scheme 4
<p>Daidzein interactions with superoxide. Bottom (σ attack): H(hydroxyl) in position 7 is captured (green arrow). When directed towards H4′ superoxide a H-bond is established. Top (π attack towards the three ring centers): when ring A is approached, the superoxide radical is later directed towards H7 (turquoise arrow), which is captured, i.e., as when originated from σ-scavenging. Interactions with rings C results in superoxide rejection, whereas when initially directed towards ring B, the superoxide behaves as its equivalent σ approach towards H4′ that is forming a H-bond.</p>
Full article ">Scheme 5
<p>Biochanin A interactions with superoxide. Bottom (σ attack): H7(hydroxyl) is scavenged (green arrow), whereas H(hydroxyl) in position 5 is rejected. Top (π attack with the three ring centers): when interacting with ring A, H7 is captured (turquoise arrow), as was the case for σ scavenging. With ring C, the formation of a π–π biochanin-η-O<sub>2</sub> complex is established, i.e., the separation between both centroids is shorter than the van der Waals distance of 3.50 Å (red arrow), 2.844 Å. With ring B, there is the rejection of superoxide.</p>
Full article ">
17 pages, 2122 KiB  
Article
Kinetics of Oxygen Reduction Reaction of Polymer-Coated MWCNT-Supported Pt-Based Electrocatalysts for High-Temperature PEM Fuel Cell
by Md Ahsanul Haque, Md Mahbubur Rahman, Faridul Islam, Abu Bakar Sulong, Loh Kee Shyuan, Ros emilia Rosli, Ashok Kumar Chakraborty and Julfikar Haider
Energies 2023, 16(3), 1537; https://doi.org/10.3390/en16031537 - 3 Feb 2023
Cited by 4 | Viewed by 2620
Abstract
Sluggish oxygen reduction reaction (ORR) of electrodes is one of the main challenges in fuel cell systems. This study explored the kinetics of the ORR reaction mechanism, which enables us to understand clearly the electrochemical activity of the electrode. In this research, electrocatalysts [...] Read more.
Sluggish oxygen reduction reaction (ORR) of electrodes is one of the main challenges in fuel cell systems. This study explored the kinetics of the ORR reaction mechanism, which enables us to understand clearly the electrochemical activity of the electrode. In this research, electrocatalysts were synthesized from platinum (Pt) catalyst with multi-walled carbon nanotubes (MWCNTs) coated by three polymers (polybenzimidazole (PBI), sulfonated tetrafluoroethylene (Nafion), and polytetrafluoroethylene (PTFE)) as the supporting materials by the polyol method while hexachloroplatinic acid (H2PtCl6) was used as a catalyst precursor. The oxygen reduction current of the synthesized electrocatalysts increased that endorsed by linear sweep voltammetry (LSV) curves while increasing the rotation rates of the disk electrode. Additionally, MWCNT-PBI-Pt was attributed to the maximum oxygen reduction current densities at −1.45 mA/cm2 while the minimum oxygen reduction current densities of MWCNT-Pt were obtained at −0.96 mAcm2. However, the ring current densities increased steadily from potential 0.6 V to 0.0 V due to their encounter with the hydrogen peroxide species generated by the oxygen reduction reactions. The kinetic limiting current densities (JK) increased gradually with the applied potential from 1.0 V to 0.0 V. It recommends that the ORR consists of a single step that refers to the first-order reaction. In addition, modified MWCNT-supported Pt electrocatalysts exhibited high electrochemically active surface areas (ECSA) at 24.31 m2/g of MWCNT-PBI-Pt, 22.48 m2/g of MWCNT-Nafion-Pt, and 20.85 m2/g of MWCNT-PTFE-Pt, compared to pristine MWCNT-Pt (17.66 m2/g). Therefore, it can be concluded that the additional ionomer phase conducting the ionic species to oxygen reduction in the catalyst layer could be favorable for the ORR reaction. Full article
(This article belongs to the Special Issue Design, Testing and Fault Diagnosis for Fuel Cells)
Show Figures

Figure 1

Figure 1
<p>FESEM images (×50 k; WD = 3 mm; EHT = 3kV) of various electrocatalysts: (<b>a</b>) MWCNT-Pt; (<b>b</b>) MWCNT-PBI-Pt; (<b>c</b>) MWCNT-Nafion-Pt; and (<b>d</b>) MWCNT-PTFE-Pt.</p>
Full article ">Figure 2
<p>TEM images of various electrocatalysts: (<b>a<sub>1</sub></b>) MWCNT-Pt; (<b>b<sub>1</sub></b>) MWCNT-PBI-Pt; (<b>c<sub>1</sub></b>) MWCNT-Nafion-Pt; and (<b>d<sub>1</sub></b>) MWCNT-PTFE-Pt. Pt nanoparticle distribution sizes of (<b>a<sub>2</sub></b>) MWCNT-Pt; (<b>b<sub>2</sub></b>) MWCNT-PBI-Pt; (<b>c<sub>2</sub></b>) MWCNT-Nafion-Pt; and (<b>d<sub>2</sub></b>) MWCNT-PTFE-Pt electrocatalysts.</p>
Full article ">Figure 3
<p>XRD graph of various electrocatalysts: (<b>a</b>) MWCNT-Pt, (<b>b</b>) MWCNT-PBI-Pt, (<b>c</b>) MWCNT-Nafion-Pt, and (<b>d</b>) MWCNT-PTFE-Pt.</p>
Full article ">Figure 4
<p>Raman spectroscopy analysis of various electrocatalysts.</p>
Full article ">Figure 5
<p>(<b>a</b>) The disk current density and (<b>b</b>) the ring current density curves of MWCNT-Nafion-Pt electrocatalyst using a rotating disk electrode loaded at various rotation rates in oxygen saturated 1.0 M H<sub>2</sub>SO<sub>4</sub> using 10 mV/s scan rate.</p>
Full article ">Figure 6
<p>Koutecky–Levich plots of various electrodes.</p>
Full article ">Figure 7
<p>The kinetic limiting current density (Jk) of various electrocatalysts at different potentials.</p>
Full article ">Figure 8
<p>(<b>a</b>) The electron transfer number (<span class="html-italic">n</span>) of various electrodes and (<b>b</b>) the percentages (%) of HO<sub>2</sub><sup>−</sup> species of various electrocatalysts.</p>
Full article ">Figure 9
<p>The linear sweep voltammetry curves of all electrodes in oxygen gas saturated 1.0 M H<sub>2</sub>SO<sub>4</sub> using a 10 mV/s scan rate and a rotation rate at 1200 rpm.</p>
Full article ">Figure 10
<p>Cyclic voltammogram (CV) of electrocatalysts in oxygen gas saturated 1.0 M H<sub>2</sub>SO<sub>4</sub> using 10 mV/s scan rate.</p>
Full article ">
13 pages, 1893 KiB  
Article
PtM/CNT (M = Mo, Ni, CoCr) Electrocatalysts with Reduced Platinum Content for Anodic Hydrogen Oxidation and Cathodic Oxygen Reduction in Alkaline Electrolytes
by Inna Vernigor, Vera Bogdanovskaya, Marina Radina, Vladimir Andreev and Oleg Grafov
Catalysts 2023, 13(1), 161; https://doi.org/10.3390/catal13010161 - 10 Jan 2023
Cited by 6 | Viewed by 1686
Abstract
Bimetallic catalysts containing platinum and transition metals (PtM, M = Mo, Ni, CoCr) were synthesized on carbon nanotubes (CNTs) functionalized in an alkaline medium. Their platinum content is 10–15% by mass. PtM/CNTNaOH are active in both the hydrogen oxidation reaction (HOR) and [...] Read more.
Bimetallic catalysts containing platinum and transition metals (PtM, M = Mo, Ni, CoCr) were synthesized on carbon nanotubes (CNTs) functionalized in an alkaline medium. Their platinum content is 10–15% by mass. PtM/CNTNaOH are active in both the hydrogen oxidation reaction (HOR) and the oxygen reduction reaction (ORR) in alkaline electrolytes. Although catalysts based on a single transition metal are inactive in the HOR, their activity in the cathode process of ORR increases relative to CNTNaOH. When using the rotating ring-disk electrode method for ORR, PtM/CNT showed a high selectivity in reducing oxygen directly to water. In HOR, the PtM/CNT catalyst had an activity comparable to that of a commercial monoplatinum catalyst. The results obtained show that it is possible to use the PtM/CNT catalyst in an alkaline fuel cell both as an anode and as a cathode. Full article
(This article belongs to the Special Issue Pt-M (M = Ni,Co,Cu, etc.)/C Electrocatalysts)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) CV for catalysts containing transition metals (shown in Figure); (<b>b</b>) CV for catalysts containing platinum (shown in Figure); 0.1 M KOH, 0.05 mV/s, m<sub>cat</sub> = 0.15 mg/cm<sup>2</sup>.</p>
Full article ">Figure 2
<p>C1s (<b>a</b>), Mo3d (<b>b</b>), O1s (<b>c</b>), and Pt4f (<b>d</b>) X-ray spectra recorded on PtMo/CNT<sub>NaOH</sub>.</p>
Full article ">Figure 3
<p>Polarization curves of hydrogen oxidation in 0.1M KOH. 0.005 V/s; 1500 rpm; m<sub>cat</sub> = 0.15 mg/cm<sup>2</sup>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Polarization curves in Tafel coordinates for platinum-modified materials (indicated in Figure), O<sub>2</sub>, 0.1 M KOH, 5 mV/s, 1500 rpm; (<b>b</b>) Chronoamperometric measurements at 0.4 V, 650 rpm; (<b>c</b>) Limiting current as a function of electrode rotation speed.</p>
Full article ">Figure 5
<p>Polarization curves on studied catalysts in 0.1 M KOH; O<sub>2</sub>; 0.005 mV/s; m<sub>cat</sub> = 0.15 mg/cm<sup>2</sup>, w = 1500 rpm.</p>
Full article ">Figure 6
<p>(<b>a</b>) Reduction polarization curves of O<sub>2</sub> on a disk electrode and their corresponding HO<sub>2</sub><sup>−</sup> oxidation curves on the ring electrode, 0.1 M KOH, 100 mcg/cm<sup>2</sup>, 5 mV/s, 1500 rpm, E<sub>ring</sub> = 1.2 V; (<b>b</b>) Dependence of the number of electrons (n) involved in the reaction (1, 2, 3, 4) and yield of hydrogen peroxide (1′, 2′, 3′, 4′) on the electrode potential.</p>
Full article ">
11 pages, 3098 KiB  
Article
Analysis of the Composition of Bromide Anion Oxidation Products in Aqueous Solutions with Different pH via Rotating Ring-Disk Electrode Method
by Roman Pichugov, Dmitry Konev, Ivan Speshilov, Lilia Abunaeva, Mikhail Petrov and Mikhail Alexeevich Vorotyntsev
Membranes 2022, 12(9), 820; https://doi.org/10.3390/membranes12090820 - 23 Aug 2022
Cited by 3 | Viewed by 1803
Abstract
We measured the ring collection coefficient of bromide anion oxidation products in a neutral and slightly alkaline medium on a rotating ring-disk electrode (glassy carbon disk, platinum ring) varying the following parameters: disk electrode rotation velocity, sodium bromide concentration, pH of the medium [...] Read more.
We measured the ring collection coefficient of bromide anion oxidation products in a neutral and slightly alkaline medium on a rotating ring-disk electrode (glassy carbon disk, platinum ring) varying the following parameters: disk electrode rotation velocity, sodium bromide concentration, pH of the medium (in the range of 6–12), anode current on the disk, and the electroreduction potential of the bromide anion oxidation products on the ring. The data obtained are presented via dependences of the cathode ring current on the disk current ratio vs. the ring electrode potential. The analysis of the results was carried out by comparing the experimental polarization curves of the ring electrode with the data of cyclic voltammetry in model solutions to determine the electrical activities of various bromine compounds in positive oxidation states. We claim that the RRDE method could be used to obtain quantitative and qualitative data on the electrooxidation of bromide ions in neutral and alkaline solutions. For the most effective regeneration of the spent oxidizer, the values of pH > 10 and moderate concentrations of NaBr should be used. Full article
(This article belongs to the Section Membrane Applications)
Show Figures

Figure 1

Figure 1
<p>Cyclic voltammograms of 0.5 M NaBr solutions measured at scan rate 50 mV/s at different pH values of buffer solutions at rotating disc and ring: (<b>a</b>) pH = 7.02 (<b>b</b>) pH = 9.93. Simultaneously, data for these solutions in the form of the ring rotating disc which collected current dependencies are presented in <a href="#membranes-12-00820-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure 2
<p>The Pt ring collected current dependencies on the potential for 0.5 M NaBr, at different pH of solutions (black and red points correspond to pH = 7.02, violet and blue points correspond to pH = 9.93) and under stationary conditions while rpm = 200. Simultaneously, data for these solutions in the form of cyclic voltammograms are presented in <a href="#membranes-12-00820-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 3
<p>Cyclic voltammograms of NaBrO<sub>3</sub> in buffer solutions with pH = 6.02 and pH = 10, recorded on a stationary platinum electrode at a scan rate of 50 mV/s.</p>
Full article ">Figure 4
<p>The Pt ring collected current vs. the potential dependence under stationary conditions for 0.1 M NaBrO<sub>3</sub> in various buffer solutions at RRDE different rotation velocities.</p>
Full article ">Figure 5
<p>Cyclic voltammograms of 0.1 M OBr<sup>−</sup> recorded at 50 mV/s on the stationary platinum electrode. pH = 11.66.</p>
Full article ">Figure 6
<p>The Pt ring collected current vs. the potential dependence under stationary conditions for 0.1 M OBr<sup>−</sup>. pH = 11.66, rpm = 200.</p>
Full article ">Figure 7
<p>Ring collection efficiencies on the Pt ring under the applied 10 mA on the GC disk for 0.5 M NaBr solutions at different pH levels. The graphs are presented for two rotation rates of the RRDE: (<b>a</b>) 200 rpm and (<b>b</b>) 5000 rpm.</p>
Full article ">Figure 8
<p>Ring collection efficiencies on the Pt ring under the applied 10 mA on the GC disk for 1 M NaBr solutions at different pH levels. The graphs are presented for two rotation rates of the RRDE: (<b>a</b>) 200 rpm and (<b>b</b>) 5000 rpm.</p>
Full article ">
21 pages, 3414 KiB  
Article
An In-Depth Exploration of the Electrochemical Oxygen Reduction Reaction (ORR) Phenomenon on Carbon-Based Catalysts in Alkaline and Acidic Mediums
by Niladri Talukder, Yudong Wang, Bharath Babu Nunna and Eon Soo Lee
Catalysts 2022, 12(7), 791; https://doi.org/10.3390/catal12070791 - 19 Jul 2022
Cited by 18 | Viewed by 6332
Abstract
Detailed studies of the electrochemical oxygen reduction reaction (ORR) on catalyst materials are crucial to improving the performance of different electrochemical energy conversion and storage systems (e.g., fuel cells and batteries), as well as numerous chemical synthesis processes. In the effort to reduce [...] Read more.
Detailed studies of the electrochemical oxygen reduction reaction (ORR) on catalyst materials are crucial to improving the performance of different electrochemical energy conversion and storage systems (e.g., fuel cells and batteries), as well as numerous chemical synthesis processes. In the effort to reduce the loading of expensive platinum group metal (PGM)-based catalysts for ORR in the electrochemical systems, many carbon-based catalysts have already shown promising results and numerous investigations on those catalysts are in progress. Most of these studies show the catalyst materials’ ORR performance as current density data obtained through the rotating disk electrode (RDE), rotating ring-disk electrode (RRDE) experiments taking cyclic voltammograms (CV) or linear sweep voltammograms (LSV) approaches. However, the provided descriptions or interpretations of those data curves are often ambiguous and recondite which can lead to an erroneous understanding of the ORR phenomenon in those specific systems and inaccurate characterization of the catalyst materials. In this paper, we presented a study of ORR on a newly developed carbon-based catalyst, the nitrogen-doped graphene/metal-organic framework (N-G/MOF), through RDE and RRDE experiments in both alkaline and acidic mediums, taking the LSV approach. The functions and crucial considerations for the different parts of the RDE/RRDE experiment such as the working electrode, reference electrode, counter electrode, electrolyte, and overall RDE/RRDE process are delineated which can serve as guidelines for the new researchers in this field. Experimentally obtained LSV curves’ shapes and their correlations with the possible ORR reaction pathways within the applied potential range are discussed in depth. We also demonstrated how the presence of hydrogen peroxide (H2O2), a possible intermediate of ORR, in the alkaline electrolyte and the concentration of acid in the acidic electrolyte can maneuver the ORR current density output in compliance with the possible ORR pathways. Full article
(This article belongs to the Special Issue Graphene in Photocatalysis/Electrocatalysis)
Show Figures

Figure 1

Figure 1
<p>LSV plot for ORR in the oxygen saturated 0.1 KOH electrolyte on the N-G/MOF (green line) and N-G (black line) catalysts coated GCE working electrode.</p>
Full article ">Figure 2
<p>LSV plots for ORR from without (green line) and with (blue line) H<sub>2</sub>O<sub>2</sub> in oxygen saturated 0.1 KOH electrolyte on the N-G/MOF coated GCE working electrode.</p>
Full article ">Figure 3
<p>LSV plot for ORR in the oxygen saturated 0.01 HClO<sub>4</sub> electrolyte on the N-G/MOF (green line) and N-G (black line) catalysts coated GC working electrode.</p>
Full article ">Figure 4
<p>LSV curves for ring-currents from the RRDE experiments with two conditions of disk surface (green and red lines) and the effect of N-G/MOF on ring-current (black line).</p>
Full article ">Figure 5
<p>LSV plot for ORR in the 0.01 M HClO<sub>4</sub> (green line), 0.005 M HClO<sub>4</sub> (blue line), and 0.001 M HClO<sub>4</sub> (red lines) electrolyte on the N-G/MOF coated GCE working electrode.</p>
Full article ">Figure 6
<p>A simplified diagram to show the function of three electrode in the RDE/RRDE system. working electrode (WE), reference electrode (RE), counter electrode (CE).</p>
Full article ">Figure 7
<p>Magnified and externally illuminated working electrode disk (GC) surface before (<b>a</b>); and after (<b>b</b>) the drop casting of N-G/MOF catalyst.</p>
Full article ">Figure 8
<p>Creating the alkaline medium ORR environment on the AFC cathode within an RDE setup: (<b>a</b>) actual test vial of our RDE setup; (<b>b</b>) schematic of working process of the H<sub>2</sub>-O<sub>2</sub> AFC; and (<b>c</b>) Schematic of the test vial of RDE setup.</p>
Full article ">Figure 9
<p>Creating the acidic medium ORR environment on the PEMFC cathode within an RDE setup: (<b>a</b>) actual test vial of our RDE setup; (<b>b</b>) schematic of working process of the H<sub>2</sub>-O<sub>2</sub> PEMFC; and (<b>c</b>) schematic of the test vial of RDE setup.</p>
Full article ">
16 pages, 3770 KiB  
Article
The Effect of an External Magnetic Field on the Electrocatalytic Activity of Heat-Treated Cyanometallate Complexes towards the Oxygen Reduction Reaction in an Alkaline Medium
by Barbara Zakrzewska, Lidia Adamczyk, Marek Marcinek and Krzysztof Miecznikowski
Materials 2022, 15(4), 1418; https://doi.org/10.3390/ma15041418 - 14 Feb 2022
Cited by 6 | Viewed by 1850
Abstract
This work focuses on the development of an electrocatalytic material by annealing a composite of a transition metal coordination material, iron hexacyanoferrate (Prussian blue) immobilized on carboxylic-acid-functionalized reduced graphene oxide. Pyrolysis at 500 °C under a nitrogen atmosphere formed nanoporous core–shell structures with [...] Read more.
This work focuses on the development of an electrocatalytic material by annealing a composite of a transition metal coordination material, iron hexacyanoferrate (Prussian blue) immobilized on carboxylic-acid-functionalized reduced graphene oxide. Pyrolysis at 500 °C under a nitrogen atmosphere formed nanoporous core–shell structures with efficient activity, which mostly included iron carbide species capable of participating in the oxygen reduction reaction in alkaline media. The physicochemical properties of the iron-based catalyst were elucidated using transmission electron microscopy, X-ray diffraction, Mössbauer spectroscopy, and various electrochemical techniques, such as cyclic voltammetry and rotating ring–disk electrode (RRDE) voltammetry. To improve the electroreduction of oxygen over the studied catalytic material, an external magnetic field was utilized, which positively shifted the potential by ca. 20 mV. The formation of undesirable intermediate peroxide species was decreased compared with the ORR measurements without an external magnetic field. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction patterns for the CFeN@rGO-COOH electrocatalysts.</p>
Full article ">Figure 2
<p>TG profile of CFeN@rGO-COOH.</p>
Full article ">Figure 3
<p>HR-TEM images of the (<b>a</b>) CFeN@rGO-COOH sample. Additionally, elementals mapping analysis of (<b>b</b>) Fe, (<b>c</b>) C, and (<b>d</b>) N are displayed.</p>
Full article ">Figure 4
<p>The Raman spectra of rGO-COOH (a) and CFeN@rGO-COOH (b).</p>
Full article ">Figure 5
<p>Room-temperature Mössbauer spectra of sample CFeN@rGO-COOH.</p>
Full article ">Figure 6
<p>Magnetic hysteresis loops of the CFeN@rGO-COOH sample.</p>
Full article ">Figure 7
<p>(<b>A</b>) Cyclic voltametric characterization of CFeN@rGO-COOH recorded in (b) O<sub>2</sub>-saturated (solid red curve) and (b’) deaerated (dash red curve) in the absence of external magnetic field. Cyclic voltametric responses of CFeN@rGO-COOH recorded in (a) O<sub>2</sub>-saturated (solid blue curve) and (a’) deaerated (dash blue curve) in the presence of external magnetic field. Insert background subtracted cathodic peak (a) with and (b) without external magnetic field. (<b>B</b>) Cyclic voltametric characterization of Pt/C nanoparticles recorded in O<sub>2</sub>-saturated (solid blue curve) and deaerated (dash red curve). Electrolyte 0.1 mol dm<sup>−3</sup> KOH. Scan rate: 50 mV s<sup>−1</sup>.</p>
Full article ">Figure 8
<p>RRDE polarization curves (background subtracted) for oxygen reduction at CFeN@rGO-COOH (a) in the absence of external magnetic field (blue curve) and (b) in the presence of external magnetic field (green curve), and (c) Pt/C nanoparticles (red curve). (<b>A</b>) Disk currents and (<b>B</b>) ring currents. Electrolyte: O<sub>2</sub>-saturated 0.1 mol dm<sup>−3</sup> KOH. Scan rate: 10 mV s<sup>−1</sup>. Rotation rate: 1600 rpm. Ring currents are obtained upon application of 1.21 V.</p>
Full article ">Figure 9
<p>Koutecky–Levich reciprocal plots for electroreduction of oxygen (at 0.75 V) at CFeN@rGO-COOH with (blue curve) and without (red curve) external magnetic field.</p>
Full article ">Figure 10
<p>(<b>A</b>) Percent fraction of peroxide anion (X% HO<sub>2</sub><sup>−</sup>) formed during electroreduction of oxygen at CFeN@rGO-COOH in (a) the absence (solid green curve) and (b) the presence (solid red curve) of an external magnetic field, and Pt/C nanoparticles (dash blue curve) under the conditions of RRDE voltametric experiment. (<b>B</b>) Numbers of transferred electrons (n) per oxygen molecule during electroreduction of oxygen under conditions described above.</p>
Full article ">
17 pages, 3738 KiB  
Article
Probing Oxygen-to-Hydrogen Peroxide Electro-Conversion at Electrocatalysts Derived from Polyaniline
by Yaovi Holade, Sarra Knani, Marie-Agnès Lacour, Julien Cambedouzou, Sophie Tingry, Teko W. Napporn and David Cornu
Polymers 2022, 14(3), 607; https://doi.org/10.3390/polym14030607 - 4 Feb 2022
Cited by 2 | Viewed by 2627
Abstract
Hydrogen peroxide (H2O2) is a key chemical for many industrial applications, yet it is primarily produced by the energy-intensive anthraquinone process. As part of the Power-to-X scenario of electrosynthesis, the controlled oxygen reduction reaction (ORR) can enable the decentralized [...] Read more.
Hydrogen peroxide (H2O2) is a key chemical for many industrial applications, yet it is primarily produced by the energy-intensive anthraquinone process. As part of the Power-to-X scenario of electrosynthesis, the controlled oxygen reduction reaction (ORR) can enable the decentralized and renewable production of H2O2. We have previously demonstrated that self-supported electrocatalytic materials derived from polyaniline by chemical oxidative polymerization have shown promising activity for the reduction of H2O to H2 in alkaline media. Herein, we interrogate whether such materials could also catalyze the electro-conversion of O2-to-H2O2 in an alkaline medium by means of a selective two-electron pathway of ORR. To probe such a hypothesis, nine sets of polyaniline-based materials were synthesized by controlling the polymerization of aniline in the presence or not of nickel (+II) and cobalt (+II), which was followed by thermal treatment under air and inert gas. The selectivity and faradaic efficiency were evaluated by complementary electroanalytical methods of rotating ring-disk electrode (RRDE) and electrolysis combined with spectrophotometry. It was found that the presence of cobalt species inhibits the performance. The selectivity towards H2O2 was 65–80% for polyaniline and nickel-modified polyaniline. The production rate was 974 ± 83, 1057 ± 64 and 1042 ± 74 µmolH2O2 h−1 for calcined polyaniline, calcined nickel-modified polyaniline and Vulcan XC 72R (state-of-the-art electrocatalyst), respectively, which corresponds to 487 ± 42, 529 ± 32 and 521 ± 37 mol kg−1cat h−1 (122 ± 10, 132 ± 8 and 130 ± 9 mol kg−1cat cm−2) for faradaic efficiencies of 58–78%. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Electrochemical characterization for the materials obtained after polymerization (5 °C, 13 h). (<b>a</b>) iR-drop uncorrected CV (N<sub>2</sub>-saturated 1 M KOH, 25 °C, 100 mV s<sup>−1</sup>, 0 rpm). (<b>b</b>) iR-drop uncorrected CV of PANI-Ni (N<sub>2</sub>-saturated 1 M KOH, 25 °C, 0 rpm) for determining ECSA. (<b>c</b>) Capacitive current (ΔI<sub>a</sub> = I<sub>a</sub>−I<sub>c</sub>) vs. scan rate at E(V<sub>RHE</sub>) = 0.8 (PANI), 1.0 (PANI-Ni) and 0.8 (PANI-Co). (<b>d</b>) iR-drop uncorrected LSV of ORR for PANI at different speeds of RRDE (O<sub>2</sub>-saturated 1 M KOH, 25 °C, 5 mV s<sup>−1</sup>, E<sub>ring</sub> = 1.2 V<sub>RHE</sub>). (<b>e</b>) iR-drop uncorrected LSV of ORR at RRDE (O<sub>2</sub>-saturated 1 M KOH, 25 °C, 5 mV s<sup>−1</sup>, 1600 rpm, E<sub>ring</sub> = 1.2 V<sub>RHE</sub>). (<b>f</b>) HO<sub>2</sub><sup>−</sup>% (left) and number of transferred electrons per molecule of O<sub>2</sub> (right <span class="html-italic">y</span>-axis) from panel (<b>e</b>).</p>
Full article ">Figure 2
<p>Electrochemical characterization for the materials obtained after polymerization (5 °C, 13 h) and stabilization (air, 350 °C, 2 h). (<b>a</b>) iR-drop uncorrected CV (N<sub>2</sub>-saturated 1 M KOH, 25 °C, 100 mV s<sup>−1</sup>, 0 rpm). (<b>b</b>) Capacitive current (ΔI<sub>a</sub> = I<sub>a</sub>−I<sub>c</sub>) vs. scan rate at E(V<sub>RHE</sub>) = 0.8 (PANI-TS), 1.2 (PANI-Ni-TS) and 0.7 (PANI-Co-TS). (<b>c</b>) iR-drop uncorrected LSV of ORR at RRDE (O<sub>2</sub>-saturated 1 M KOH, 25 °C, 5 mV s<sup>−1</sup>, 1600 rpm, E<sub>ring</sub> = 1.2 V<sub>RHE</sub>). (<b>d</b>) HO<sub>2</sub><sup>−</sup>% (left) and number of transferred electrons per molecule of O<sub>2</sub> (right <span class="html-italic">y</span>-axis) from panel (<b>c</b>).</p>
Full article ">Figure 3
<p>Electrochemical characterization for the materials obtained after polymerization (5 °C, 13 h), stabilization (air, 350 °C, 2 h) and calcination (N<sub>2</sub>, 900 °C, 6 h). (<b>a</b>) iR-drop uncorrected CV (N<sub>2</sub>-saturated 1 M KOH, 25 °C, 100 mV s<sup>−1</sup>, 0 rpm). (<b>b</b>) iR-drop uncorrected CV of PANI-TS-TC (N<sub>2</sub>-saturated 1 M KOH, 25 °C, 0 rpm) for determining ECSA. (<b>c</b>) Capacitive current (ΔI<sub>a</sub> = I<sub>a</sub>−I<sub>c</sub>) vs. scan rate at E(V<sub>RHE</sub>) = 0.8 (PANI-TS-TC), 0.9 (PANI-Ni-TS-TC), 0.7 (PANI-Co-TS-TC) and 1.0 (Vulcan). (<b>d</b>) iR-drop uncorrected LSV of ORR for PANI-TS-TC at different speeds of RRDE (O<sub>2</sub>-saturated 1 M KOH, 25 °C, 5 mV s<sup>−1</sup>, E<sub>ring</sub> = 1.2 V<sub>RHE</sub>). (<b>e</b>) iR-drop uncorrected LSV of ORR at RRDE (O<sub>2</sub>-saturated 1 M KOH, 25 °C, 5 mV s<sup>−1</sup>, 1600 rpm, E<sub>ring</sub> = 1.2 V<sub>RHE</sub>). (<b>f</b>) HO<sub>2</sub><sup>−</sup>% (left) and number of transferred electrons per molecule of O<sub>2</sub> (right <span class="html-italic">y</span>-axis) from panel (<b>e</b>).</p>
Full article ">Figure 4
<p>Bulk electrolysis and electroanalysis. (<b>a</b>) Electrical charge passed during the potentiostatic electrolysis at an applied potential of 0.6 V<sub>RHE</sub> (iR-uncorrected). (<b>b</b>) UV-vis spectra of the anodic compartment at different times of the electrolysis (aliquot was neutralized by sulfuric acid before addition of potassium titanium oxalate solution). (<b>c</b>) UV-vis spectrum of hydrogen peroxide (638 µM) in the presence of sulfuric and potassium titanium oxalate solution: inset shows the color of the solution in the presence of increasing concentration of hydrogen peroxide. (<b>d</b>) Calibration curves from UV-vis assays at 390 nm and from solutions of panel (<b>c</b>). (<b>e</b>) Hydrogen peroxide productivity as quantified by the UV-vis assays. (<b>f</b>) Faradaic efficiency. The electrocatalyst loading was 0.5 mg cm<sup>−2</sup> and the electrode size was 4 cm<sup>2</sup>. The blank electrode was bare carbon paper MGL370 (Fuel Cell Earth LLC). Error bars represent 1 SD (n ≥ 3).</p>
Full article ">Figure 5
<p>SEM characterization of the materials obtained after polymerization (5 °C, 13 h), stabilization (air, 350 °C, 2 h) and calcination (N<sub>2</sub>, 900 °C, 6 h): backscattered SEM images at different magnification for: (<b>a1</b>,<b>a2</b>) PANI-TS-TC, (<b>b1</b>,<b>b2</b>) PANI-Ni-TS-TC, and (<b>c1</b>,<b>c2</b>) PANI-Co-TS-TC.</p>
Full article ">Figure 6
<p>Materials obtained after polymerization (5 °C, 13 h), stabilization (air, 350 °C, 2 h) and calcination (N<sub>2</sub>, 900 °C, 6 h): EDX spectra, backscattered SEM images plus EDX maps of: (<b>a</b>) PANI-Ni-TS-TC, and (<b>b</b>) PANI-Co-TS-TC.</p>
Full article ">Figure 7
<p>XRD patterns the electrocatalysts obtained after polymerization (5 °C, 13 h), stabilization (air, 350 °C, 2 h) and calcination (N<sub>2</sub>, 900 °C, 6 h): (<b>a</b>) PANI-TS-TC and PANI-Ni-TS-TC, and (<b>b</b>) PANI-Co-TS-TC.</p>
Full article ">
11 pages, 2612 KiB  
Article
Ultrafine TaOx/CB Oxygen Reduction Electrocatalyst Operating in Both Acidic and Alkaline Media
by Jun-Woo Park and Jeongsuk Seo
Catalysts 2022, 12(1), 35; https://doi.org/10.3390/catal12010035 - 29 Dec 2021
Cited by 6 | Viewed by 1820
Abstract
The high activity of non-platinum electrocatalysts for oxygen reduction reaction (ORR) in alkaline media is necessary for applications in energy conversion devices such as fuel cells and metal-air batteries. Herein, we present the electrocatalytic activity of TaOx/carbon black (CB) nanoparticles for [...] Read more.
The high activity of non-platinum electrocatalysts for oxygen reduction reaction (ORR) in alkaline media is necessary for applications in energy conversion devices such as fuel cells and metal-air batteries. Herein, we present the electrocatalytic activity of TaOx/carbon black (CB) nanoparticles for the ORR in an alkaline atmosphere as well as in an acidic electrolyte. Ultrafine TaOx nanoparticles 1–2 nm in size and uniformly dispersed on CB supports were prepared by potentiostatic electrodeposition in a nonaqueous electrolyte and subsequent annealing treatment in an H2 flow. The TaOx/CB nanoparticles largely catalyzed the ORR with an onset potential of 1.03 VRHE in an O2-saturated 0.1 M KOH solution comparable to that of a commercial Pt/CB catalyst. ORR activity was also observed in 0.1 M H2SO4 solution. According to the rotating ring disk electrode measurement results, the oxide nanoparticles partly produced H2O2 during the ORR in 0.1 M KOH, and the ORR process was dominated by both the two- and four-electron reductions of oxygen in a diffusion-limited potential region. The Tafel slope of −120 mV dec−1 in low and high current densities revealed the surface stability of the oxide nanoparticles during the ORR. Therefore, these results demonstrated that the TaOx/CB nanoparticles were electroactive for the ORR in both acidic and alkaline electrolytes. Full article
Show Figures

Figure 1

Figure 1
<p>Surface morphology of ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles (<b>a</b>) TEM and (<b>b</b>) STEM images of ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow. (<b>c</b>) Size distribution and (<b>d</b>) elemental EDS mapping results for the prepared fine particles shown in (<b>b</b>).</p>
Full article ">Figure 2
<p>Narrow-scan (<b>a</b>) Ta 4f and (<b>b</b>) O 1s XPS spectra of ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow.</p>
Full article ">Figure 3
<p>CVs of the ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow. The electrochemical measurements were carried out in Ar- and O<sub>2</sub>-saturated 0.1 M KOH aqueous solutions at a scan rate of 100 mV s<sup>−1</sup>.</p>
Full article ">Figure 4
<p>ORR activity of ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles (<b>a</b>) LSV of the ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow for ORR in O<sub>2</sub>-purged 0.1 M KOH aqueous solution at a rotation rate of 1600 rpm. For comparison, commercial 10 and 20 wt% Pt/CB catalysts were also measured in the same manner. (<b>b</b>) The LSVs of the prepared TaO<span class="html-italic"><sub>x</sub></span>/CB and bare CB nanoparticles for ORR in O<sub>2</sub>-purged 0.1 M KOH (dotted lines) and 0.1 M H<sub>2</sub>SO<sub>4</sub> aqueous solutions (solid lines). (<b>c</b>) Tafel plots for TaO<span class="html-italic"><sub>x</sub></span>/CB, bare CB, and commercial 20 wt% Pt/CB catalysts, determined from the LSVs measured in O<sub>2</sub>-saturated 0.1 M KOH aqueous solutions. (<b>d</b>) A chronoamperometry curve of the TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles applied at 0.7 V<sub>RHE</sub> in O<sub>2</sub>-saturated 0.1 M KOH aqueous electrolyte for 5 h.</p>
Full article ">Figure 5
<p>LSVs of the RRDE measurements for ORR over ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow in O<sub>2</sub>-purged 0.1 M KOH and 0.1 M H<sub>2</sub>SO<sub>4</sub> aqueous solutions at a revolution rate of 1600 rpm and a sweep rate of 5 mV s<sup>−1</sup>. A potential of 1.2 V<sub>RHE</sub> was applied to the Pt ring electrode.</p>
Full article ">Figure 6
<p>Hydrodynamic voltammetry of ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles (<b>a</b>) LSVs of the ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles prepared by electrodeposition and subsequent annealing treatment in an H<sub>2</sub> flow in O<sub>2</sub>-purged 0.1 M KOH aqueous solutions at a sweep rate of 5 mV s<sup>−1</sup> obtained at different revolution rates of 100, 400, 900, 1600, and 2500 rpm. (<b>b</b>) Koutecky–Levich plots, |<span class="html-italic">i|</span><sup>−1</sup> versus <span class="html-italic">ω</span><sup>−1/2</sup>, for the ultrafine TaO<span class="html-italic"><sub>x</sub></span>/CB nanoparticles calculated from the LSVs in (<b>a</b>). The dotted lines indicate the slopes determined from the two- and four-electron mechanisms.</p>
Full article ">
Back to TopTop