[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,298)

Search Parameters:
Keywords = transition metal oxides

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 5118 KiB  
Article
Enhanced Performance of Sodium-Ion Battery Cathodes with Ti and V Co-Doped P2-Type Na0.67Fe0.5Mn0.5O2 Materials
by Trapa Banik, Indranil Bhattacharya, Kirankumar Venkatesan Savunthari, Sanjeev Mukerjee, Webster Adepoju and Abiodun Olatunji
Electrochem 2024, 5(4), 437-454; https://doi.org/10.3390/electrochem5040029 - 18 Oct 2024
Viewed by 142
Abstract
Manganese- and iron-rich P2-type Na0.67Fe0.5Mn0.5O2 (NFM) has garnered significant interest as a promising cathode candidate due to the natural abundance of Fe and Mn along with a high [...] Read more.
Manganese- and iron-rich P2-type Na0.67Fe0.5Mn0.5O2 (NFM) has garnered significant interest as a promising cathode candidate due to the natural abundance of Fe and Mn along with a high redox couple of Fe3+/Fe4+ and Mn3+/Mn4+. Despite all these merits, NFM suffers from structural instability during cycling, arising from the destructive Jahn-Teller (JT) distortion effect of Mn3+/Mn4+ during charging and Fe4+/Fe3+ during discharging. In this research, a novel P2-type transition metal-oxide cathode Na0.67Fe0.5−2xMn0.5TixVxO2 was synthesized by doping a tiny fraction of two electrochemically inactive elements, Titanium (Ti) and Vanadium (V), into Mn-rich Na0.67Fe0.5Mn0.5O2 (NFM) that mitigated the JT effect substantially and ameliorated the stability of the SIB during cycling. These exhaustive structural and morphological comparisons provided insights into the effects of V and Ti doping on stabilizing surface structures, reducing Jahn Teller distortion, enhancing stability and capacity retention, and promoting the Na+ carrier transport mechanism. Moreover, the electrochemical analysis, such as the galvanostatic charge/discharge profile, validates the capacity improvement via Ti and V co-doping into NFM cathode. The initial discharge capacity of the 2% Ti/V-doped Na0.67Fe0.48Mn0.5Ti0.01V0.01O2 (2NFMTV) was found to be 187.12 mAh g−1 at a rate of 0.1 C, which was greater than the discharge capacity of 175.15 mAh g−1 observed for pure NFM (Na0.67Mn0.5Fe0.5O2). In contrast, 2NFMTV exhibited a noteworthy capacity retention of 46.1% when evaluated for its original capacity after undergoing 150 cycles at a rate of 0.1 C. This research also established a structural doping approach as a feasible technique for advancing the progress of next-generation Sodium-ion Batteries. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of (<b>a</b>) pristine NFM and Ti/V-doped NaFe<sub>0.5−2x</sub>Mn<sub>0.5</sub>Ti<sub>x</sub>V<sub>x</sub>O<sub>2</sub> (x = 0~0.1), (<b>b</b>) lattice structure of 2NFMTV, (<b>c</b>) lattice parameters obtained from Rietveld refinement, and (<b>d</b>) schematic illustration of P2-type layered transition metal oxides.</p>
Full article ">Figure 2
<p>Rietveld refinement on XRD data of (<b>a</b>) pristine NFM, (<b>b</b>) 1NFMTV, (<b>c</b>) 2NFMTV, and (<b>d</b>) 3NFMTV (blue lines and red lines correspond to observed intensities and differences between observed and calculated patterns, respectively).</p>
Full article ">Figure 3
<p>SEM and TEM images of (<b>a</b>) pure NFM (<b>b</b>) 2NFMTV.</p>
Full article ">Figure 4
<p>Elemental mapping from EDS Analysis of (<b>a</b>) pure NFM (<b>b</b>) 2NFMTV.</p>
Full article ">Figure 5
<p>XPS analysis of all elements in pristine NFM and Ti/V-doped NFM (1 NFMTV, 2NFMTV, 3NFMTV). (<b>a</b>) Na 1s Peak; (<b>b</b>) Fe 2s Peaks; (<b>c</b>) Mn 2p Peaks; (<b>d</b>) Mn 3s Peaks; (<b>e</b>) Ti 2p Peaks; (<b>f</b>) O 1s and V 2p Peaks.</p>
Full article ">Figure 6
<p>(<b>a</b>) Raman Spectroscopy and (<b>b</b>) FTIR analysis of pristine and Ti/V-dopedNFM samples.</p>
Full article ">Figure 7
<p>Galvanostatic charge/discharge data for (<b>a</b>) pure NFM (<b>b</b>) 2NFMTV (<b>c</b>) long-term cycling stability tests for pure NFM and 2NFMTV and (<b>d</b>) rate capability of pure NFM and 2NFMTV.</p>
Full article ">Figure 8
<p>(<b>a</b>) CV for pure NFM and 2NFMTV (<b>b</b>) linear response of peak current as a function of the square root of scan rate of 2NFMTV, (<b>c</b>) CV for 2NFMTV at various scan rates of 0.1–5 mV s<sup>−1</sup>, (<b>d</b>) Nyquist plots over a frequency range from 0.1 kHz to 100 kHz and equivalent circuit of pure NFM and 2NFMTV.</p>
Full article ">
33 pages, 3665 KiB  
Review
Role of Sintering Aids in Electrical and Material Properties of Yttrium- and Cerium-Doped Barium Zirconate Electrolytes
by Shivesh Loganathan, Saheli Biswas, Gurpreet Kaur and Sarbjit Giddey
Processes 2024, 12(10), 2278; https://doi.org/10.3390/pr12102278 - 18 Oct 2024
Viewed by 211
Abstract
Ceramic proton conductors have the potential to lower the operating temperature of solid oxide cells (SOCs) to the intermediate temperature range of 400–600 °C. This is attributed to their superior ionic conductivity compared to oxide ion conductors under these conditions. However, prominent proton-conducting [...] Read more.
Ceramic proton conductors have the potential to lower the operating temperature of solid oxide cells (SOCs) to the intermediate temperature range of 400–600 °C. This is attributed to their superior ionic conductivity compared to oxide ion conductors under these conditions. However, prominent proton-conducting materials, such as yttrium-doped barium cerates and zirconates with specified compositions like BaCe1−xYxO3−δ (BCY), BaZr1−xYxO3−δ (BZY), and Ba(Ce,Zr)1−yYyO3−δ (BCZY), face significant challenges in achieving dense electrolyte membranes. It is suggested that the incorporation of transition and alkali metal oxides as sintering additives can induce liquid phase sintering (LPS), offering an efficient method to facilitate the densification of these proton-conducting ceramics. However, current research underscores that incorporating these sintering additives may lead to adverse secondary effects on the ionic transport properties of these materials since the concentration and mobility of protonic defects in a perovskite are highly sensitive to symmetry change. Such a drop in ionic conductivity, specifically proton transference, can adversely affect the overall performance of cells. The extent of variation in the proton conductivity of the perovskite BCZY depends on the type and concentration of the sintering aid, the nature of the sintering aid precursors used, the incorporation technique, and the sintering profile. This review provides a synopsis of various potential sintering techniques, explores the influence of diverse sintering additives, and evaluates their effects on the densification, ionic transport, and electrochemical properties of BCZY. We also report the performance of most of these combinations in an actual test environment (fuel cell or electrolysis mode) and comparison with BCZY. Full article
(This article belongs to the Section Chemical Processes and Systems)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Schematic showing working principle of proton-conducting solid oxide fuel cell (H-SOFCs) (<b>a</b>) and proton-conducting solid oxide electrolytic cell (H-SOECs) (<b>b</b>).</p>
Full article ">Figure 2
<p>Depiction of bulk, grain, and electrode resistances in Nyquist plot for BCZY (<b>a</b>), reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>]; BCZY electrolyte film thickness and porosity as function of sintering temperature replotted [<a href="#B44-processes-12-02278" class="html-bibr">44</a>] (<b>b</b>); X-ray diffractogram showing peak shift of BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> (BCZY35) upon adding NiO, CuO, and ZnO sintering aids, reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>] (<b>c</b>).</p>
Full article ">Figure 3
<p>Arrhenius plots of total (<b>a</b>), grain boundary (<b>b</b>), and grain bulk (<b>c</b>) conductivity of 1 wt% NiO-, ZnO-, and CuO-added BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> (BCZY35) in 3% humid H<sub>2</sub> reproduced with permission from Elsevier [<a href="#B54-processes-12-02278" class="html-bibr">54</a>]; Arrhenius plots of grain interior (GI), grain boundary (GB), and total conductivity of BZY10 + 0.2 wt%NiO sintered at 1500 (<b>d</b>) and BZY10 sintered at 1600 (<b>e</b>) under 2.7% humidified H<sub>2</sub> and Ar (1:19 <span class="html-italic">v</span>/<span class="html-italic">v</span>), replotted from [<a href="#B95-processes-12-02278" class="html-bibr">95</a>].</p>
Full article ">Figure 4
<p>Cross-sectional FE-SEM images for BCZY63 pellets doped with (<b>A</b>) CuO—1450 °C, (<b>B</b>) ZnO—1450 °C, (<b>C</b>) Fe<sub>2</sub>O<sub>3</sub>—1600 °C, (<b>D</b>) Cr<sub>2</sub>O<sub>3</sub>—1600 °C, (<b>E</b>) PdO—1600 °C, and (<b>F</b>) Control—1650 °C, reproduced with permission from Elsevier [<a href="#B109-processes-12-02278" class="html-bibr">109</a>]. Protonic conductivity as a function of temperature for BCZY doped with ZnO and CuO (<b>G</b>), Fe<sub>2</sub>O<sub>3</sub> and MnO<sub>2</sub> (<b>H</b>), and Cr<sub>2</sub>O<sub>3</sub> and PdO (<b>I</b>) under different atmospheres, replotted from an original paper [<a href="#B109-processes-12-02278" class="html-bibr">109</a>]. The schematic shows that no sintering aid leads to porosity, whereas an excess of one causes the accumulation of unwanted secondary phases along the grain boundaries (<b>J</b>).</p>
Full article ">Figure 5
<p>Relative density of Ba<sub>1.03</sub>Ce<sub>0.5</sub>Zr<sub>0.4</sub>Y<sub>0.1</sub>O<sub>3−<span class="html-italic">δ</span></sub> with various ZnO concentrations sintered at different temperatures (<b>a</b>) and SEM images of Ba<sub>1.03</sub>Ce<sub>0.5</sub>Zr<sub>0.4</sub>Y<sub>0.1</sub>O<sub>3−<span class="html-italic">δ</span></sub> without (<b>b</b>) and with 1 wt% (<b>c</b>) ZnO sintered at 1300 °C for 10 h, reproduced with permission from Elsevier [<a href="#B115-processes-12-02278" class="html-bibr">115</a>]; BaCe<sub>0.8</sub>Zr<sub>0.1</sub>Y<sub>0.1</sub>O<sub>3−δ</sub> densification mechanism through ZnO.BaO eutectic formation reproduced with permission from Elsevier [<a href="#B116-processes-12-02278" class="html-bibr">116</a>] (<b>d</b>); thermogravimetric analysis (TGA) curve of pristine BaCe<sub>0.3</sub>Zr<sub>0.55</sub>Y<sub>0.15</sub>O<sub>3−δ</sub> [<a href="#B54-processes-12-02278" class="html-bibr">54</a>] in pure CO<sub>2</sub> environment (<b>e</b>), and TGA curve of BaCe<sub>0.55</sub>Zr<sub>0.3</sub>Y<sub>0.15</sub>O<sub>3−d</sub> and BaCe<sub>0.35</sub>Zr<sub>0.5</sub>Y<sub>0.15</sub>O<sub>3−d</sub> in CO<sub>2</sub> and air (<b>f</b>) [<a href="#B121-processes-12-02278" class="html-bibr">121</a>], replotted from original papers.</p>
Full article ">Figure 6
<p>The temperature dependence of the conductivity of different BCZY10 and BCZY27 samples under a wet reducing atmosphere (9% H<sub>2</sub>N<sub>2</sub>, P<sub>H2O</sub> = 0.015 atm) with (<b>a</b>) Co doping and (<b>b</b>) Ni doping [<a href="#B132-processes-12-02278" class="html-bibr">132</a>]; Arrhenius plots of the conductivities in 3% H<sub>2</sub>O of the BZCY, BZCY-2, BZCY-5, and BZCY-10 samples sintered at 1400 °C for 5 h, reproduced with permission from Elsevier [<a href="#B152-processes-12-02278" class="html-bibr">152</a>] (<b>c</b>).</p>
Full article ">Figure 7
<p>Various strategies that require further in-depth investigation to achieve BCZY densification at lower temperatures using sintering aids without any deterioration of protonic conductivity.</p>
Full article ">
17 pages, 1606 KiB  
Article
Dopaminergic- and Serotonergic-Dependent Behaviors Are Altered by Lanthanide Series Metals in Caenorhabditis elegans
by Anthony Radzimirski, Michael Croft, Nicholas Ireland, Lydia Miller, Jennifer Newell-Caito and Samuel Caito
Toxics 2024, 12(10), 754; https://doi.org/10.3390/toxics12100754 - 17 Oct 2024
Viewed by 265
Abstract
The lanthanide series elements are transition metals used as critical components of electronics, as well as rechargeable batteries, fertilizers, antimicrobials, contrast agents for medical imaging, and diesel fuel additives. With the surge in their utilization, lanthanide metals are being found more in our [...] Read more.
The lanthanide series elements are transition metals used as critical components of electronics, as well as rechargeable batteries, fertilizers, antimicrobials, contrast agents for medical imaging, and diesel fuel additives. With the surge in their utilization, lanthanide metals are being found more in our environment. However, little is known about the health effects associated with lanthanide exposure. Epidemiological studies as well as studies performed in rodents exposed to lanthanum (La) suggest neurological damage, learning and memory impairment, and disruption of neurotransmitter signaling, particularly in serotonin and dopamine pathways. Unfortunately, little is known about the neurological effects of heavier lanthanides. As dysfunctions of serotonergic and dopaminergic signaling are implicated in multiple neurological conditions, including Parkinson’s disease, depression, generalized anxiety disorder, and post-traumatic stress disorder, it is of utmost importance to determine the effects of La and other lanthanides on these neurotransmitter systems. We therefore hypothesized that early-life exposure of light [La (III) or cerium (Ce (III))] or heavy [erbium (Er (III)) or ytterbium (Yb (III))] lanthanides in Caenorhabditis elegans could cause dysregulation of serotonergic and dopaminergic signaling upon adulthood. Serotonergic signaling was assessed by measuring pharyngeal pump rate, crawl-to-swim transition, as well as egg-laying behaviors. Dopaminergic signaling was assessed by measuring locomotor rate and egg-laying and swim-to-crawl transition behaviors. Treatment with La (III), Ce (III), Er (III), or Yb (III) caused deficits in serotonergic or dopaminergic signaling in all assays, suggesting both the heavy and light lanthanides disrupt these neurotransmitter systems. Concomitant with dysregulation of neurotransmission, all four lanthanides increased reactive oxygen species (ROS) generation and decreased glutathione and ATP levels. This suggests increased oxidative stress, which is a known modifier of neurotransmission. Altogether, our data suggest that both heavy and light lanthanide series elements disrupt serotonergic and dopaminergic signaling and may affect the development or pharmacological management of related neurological conditions. Full article
(This article belongs to the Special Issue Heavy Metal Induced Neurotoxicity)
Show Figures

Figure 1

Figure 1
<p>Lanthanide-induced toxicity in <span class="html-italic">C. elegans</span>. L1-stage N2 worms were treated with increasing concentrations of (<b>top-left</b>) LaCl<sub>3</sub>, (<b>top-right</b>) CeCl<sub>2</sub>, (<b>bottom-left</b>) ErCl<sub>3</sub>, or (<b>bottom-right</b>) YbCl<sub>3</sub> for 30 min, washed, and transferred to agar plate supplemented with OP50 <span class="html-italic">Escherichia coli</span>. Survival was manually counted 48 h post-exposure and fit to a non-linear sigmoidal dose response. LD50 values were calculated and reported. Data points represent the mean + SEM from 5 independent experiments.</p>
Full article ">Figure 2
<p>Lanthanides decrease DAergic function in <span class="html-italic">C. elegans</span>. L1-stage N2 worms were treated with lanthanides for 30 min, washed, and maintained on NGM agar plates seeded with OP50 <span class="html-italic">E.</span> coli. A total of 72 h post-exposure, worms were analyzed for the (<b>A</b>) BSR or (<b>B</b>) swim-to-crawl transition. (<b>C</b>) Alternatively, 30 h after reaching the L4 stage, eggs were quantified on agar or unlaid inside the worm. Data are presented as average (<b>A</b>) change in body bends on vs. off <span class="html-italic">E. coli</span>, (<b>B</b>) time to crawl, or (<b>C</b>) % eggs laid ± SEM of five independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. # <span class="html-italic">p</span> &lt; 0.05 and ### <span class="html-italic">p</span> &lt; 0.001 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 3
<p>Lanthanides decrease serotonergic function in <span class="html-italic">C. elegans</span>. L1-stage (<b>A</b>) PE254 or (<b>B</b>) N2 worms were treated with lanthanides for 30 min, washed, and maintained on NGM agar plates seeded with OP50 <span class="html-italic">E. coli</span>. A total of 72 h post-exposure, worms were analyzed for (<b>A</b>) pharynx pump rate or (<b>B</b>) crawl-to-swim transition. Data are presented as average (<b>A</b>) rate of luminescence for 1 h and (<b>B</b>) time to swim ± SEM of five independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. ## <span class="html-italic">p</span> &lt; 0.01 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 4
<p>Lanthanides decrease dopamine and serotonin in <span class="html-italic">C. elegans</span>. (<b>A</b>) DA or (<b>B</b>) 5-HT was quantified immediately following exposure to La (III), Ce (III), Er (III), or Yb (III). Data are expressed as means ± SEM of five independent experiments. ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 as compared to untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. ## <span class="html-italic">p</span> &lt; 0.01 and ### <span class="html-italic">p</span> &lt; 0.001 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">Figure 5
<p>Lanthanides induce oxidative stress in <span class="html-italic">C. elegans</span>. Worms were treated for 30 min with La (III), Ce (III), Er (III), or Yb (III). Measures of oxidative stress were either assessed immediately (<b>A</b>) or 24 h following exposure (<b>B</b>–<b>E</b>). (<b>A</b>) ROS levels were assessed by DCFDA fluorescence. Data are expressed as mean fluorescence ± SEM for five independent experiments (<b>B</b>) AGE protein adducts were measured and normalized to protein content. Data are expressed as means ± SEM from 5 independent experiments. (<b>C</b>) Total GSH levels were quantified and normalized to protein content. Data are expressed as mean ± SEM from five independent experiments. (<b>D</b>) Quantification of GFP fluorescence of VP596 transgenic worms. Data are expressed as mean fluorescence ± SEM from 5 independent experiments. (<b>E</b>) ATP levels were quantified and normalized to protein content. Data are expressed as mean ± SEM from five independent experiments. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001 as compared with untreated control. Horizontal bars represent comparisons between concentrations of lanthanide-treated worms. # <span class="html-italic">p</span> &lt; 0.05, ## <span class="html-italic">p</span> &lt; 0.01, and ### <span class="html-italic">p</span> &lt; 0.001 as compared between lower and higher concentrations of lanthanides.</p>
Full article ">
10 pages, 1029 KiB  
Article
Theoretical Investigation of Electric Polarizability in Porphyrin–Zinc and Porphyrin–Zinc–Thiazole Complexes Using Small Property-Oriented Basis Sets
by Arkadiusz Kuziemski, Krzysztof Z. Łączkowski and Angelika Baranowska-Łączkowska
Int. J. Mol. Sci. 2024, 25(20), 11044; https://doi.org/10.3390/ijms252011044 - 14 Oct 2024
Viewed by 313
Abstract
Porphyrin complexes are of great importance due to their possible applications as sensors, solar cells and photocatalysts, as well as their ability to bind additional ligands. A valuable source of knowledge on their nature is their electric properties, which can be evaluated employing [...] Read more.
Porphyrin complexes are of great importance due to their possible applications as sensors, solar cells and photocatalysts, as well as their ability to bind additional ligands. A valuable source of knowledge on their nature is their electric properties, which can be evaluated employing density functional theory (DFT) methods, supporting the experimental research. The present work aims at the application of small property-oriented basis sets in calculation of electric properties in transition metals, their oxides and test coordination complexes. Firstly, the existing polarized ZPol basis set for the first-row transition metals is modified in order to improve atomic polarizability results. For this purpose, optimization of the f-type polarization function exponent is carried out with respect to the value of average atomic polarizability of investigated metals. Next, both the original and the modified basis sets are employed in finite field CCSD(T) calculation of transition metal oxides’ dipole moments, as well as DFT calculation of polarizabilities in porphyrin–zinc and porphyrin–zinc–thiazole complexes. The obtained results show that the ZPol and ZPol-A basis sets can be successfully employed in the calculation of linear electric properties in large systems. The optimization procedure used in the present work can be employed for other source basis sets and elements, leading to new efficient polarized basis sets. Full article
(This article belongs to the Special Issue Molecular Modeling: Latest Advances and Applications)
Show Figures

Figure 1

Figure 1
<p>The B3LYP/6-311G**/LanL2DZ optimized structure of the porphyrin–zinc complex. White, light grey, dark grey and blue spheres represent hydrogen, carbon, zinc and nitrogen atoms, respectively. Values of geometrical parameters are reported in <a href="#app1-ijms-25-11044" class="html-app">Table S1 of Supporting Information</a>.</p>
Full article ">Figure 2
<p>The B3LYP/6-311G**/LanL2DZ optimized structure of the porphyrin–zinc–thiazol complex. White, light grey, dark grey, yellow and blue spheres represent hydrogen, carbon, zinc, sulfur and nitrogen atoms, respectively. Values of geometrical parameters are reported in <a href="#app1-ijms-25-11044" class="html-app">Table S2 of Supporting Information</a>.</p>
Full article ">
16 pages, 2753 KiB  
Article
Hydrogenation Studies of Iridium Pyridine Diimine Complexes with O- and S-Donor Ligands (Hydroxido, Methoxido and Thiolato)
by Max Völker, Matthias Schreyer and Peter Burger
Chemistry 2024, 6(5), 1230-1245; https://doi.org/10.3390/chemistry6050071 - 11 Oct 2024
Viewed by 432
Abstract
For square-planar late transition metal pyridine, diimine (Rh, Ir) complexes with hydro-xido, methoxido, and thiolato ligands. We could previously establish sizable metal-O- and S π-bonding interactions. Herein, we report the hydrogenation studies of iridium hydroxido and methoxido complexes, which quantitatively lead to the [...] Read more.
For square-planar late transition metal pyridine, diimine (Rh, Ir) complexes with hydro-xido, methoxido, and thiolato ligands. We could previously establish sizable metal-O- and S π-bonding interactions. Herein, we report the hydrogenation studies of iridium hydroxido and methoxido complexes, which quantitatively lead to the trihydride compound and water/methanol. The iridium trihydride displays a highly fluctional structure with scrambling hydrogen atoms, which can be described as a dihydrogen hydride system based on NMR and DFT investigations. This contrasts the iridium sulfur compounds, which are not reacting with dihydrogen. According to DFT and LNO-CCSD(T) calculations, hydrogenation of the methoxido complex proceeds by a two-step mechanism, i.e., an oxidative addition step of H2 to an Ir(III) dihydride intermediate with consecutive reductive O-H elimination of methanol. Based on PNO-CCSD(T) calculations, the reactivity difference between the O- and S-donors can be traced to the stronger H-O bonds in the water/methanol products compared to the S-H bonds in the sulphur congeners, which serves as a driving force for hydrogenation. Full article
(This article belongs to the Section Inorganic and Solid State Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>vT <sup>1</sup>H-NMR spectrum of <b>2</b> in toluene-d<sub>8</sub> in the temperature range of 176–296 K. The hydride resonance is located at −9.5 ppm.</p>
Full article ">Figure 2
<p>X-ray crystal structure of the trihydride complex <b>7</b> (Ortep plot at the 50% probability level).</p>
Full article ">Figure 3
<p>Geometry optimized structure of the trihydride complex <b>7</b> with selected distances in Å. The phenyl substituents of the imine carbon and 2,6-aryl group of the N<sub>imine</sub> atoms are omitted for clarity.</p>
Full article ">Figure 4
<p>Contour plot of the energy hypersurface for the energy dependence (in [kcal/mol]) on the H<sub>apical</sub>-H<sub>equatorial</sub> distances r1 and r2 in [Å] with highlighted values of contour lines.</p>
Full article ">Figure 5
<p>Transition states (<b>left</b> and <b>right</b>) for hydrogen scrambling in the trihydrido complex <b>7</b>. Arrows indicate the corresponding mode. Selected distances in Å are presented.</p>
Full article ">Figure 6
<p>Mechanism of the hydrogenation process in the methoxido complex <b>4</b>. Relative energies are given in kcal/mol (LNO-CCSD(T)/def2-TZVPP//DFT(PBE-D3BJ, def2-TZVP)). The aryl groups of the ketimine units were omitted for clarity. For the full structures and optimized coordinates (see <a href="#app1-chemistry-06-00071" class="html-app">SI</a>).</p>
Full article ">Scheme 1
<p>Reprinted with permission from Goldberg et al. [<a href="#B1-chemistry-06-00071" class="html-bibr">1</a>] <span class="html-italic">J. Am. Chem. Soc.</span> <b>2011</b>, 133, 44. Copyright 2011 American Chemical Society.</p>
Full article ">Scheme 2
<p>Hydrogenation of the iridium methyl compound <b>1</b> to the trihydride complex <b>2</b>.</p>
Full article ">Scheme 3
<p>Hydrogenation reaction of the PDI complexes <b>3</b>–<b>6</b> with O and S donor ligands.</p>
Full article ">Scheme 4
<p>Mechanism of the hydrogen atom scrambling.</p>
Full article ">
14 pages, 7065 KiB  
Article
Sustainable Synthesis of a Carbon-Supported Magnetite Nanocomposite Anode Material for Lithium-Ion Batteries
by Hui Zeng, Jiahui Li, Haoyu Yin, Ruixin Jia, Longbiao Yu, Hongliang Li and Binghui Xu
Batteries 2024, 10(10), 357; https://doi.org/10.3390/batteries10100357 - 11 Oct 2024
Viewed by 485
Abstract
Transition metal oxide magnetite (Fe3O4) is recognized as a potential anode material for lithium-ion batteries owing to its high theoretical specific capacity, modest voltage output, and eco-friendly character. It is a challenging task to engineer high-performance composite materials by [...] Read more.
Transition metal oxide magnetite (Fe3O4) is recognized as a potential anode material for lithium-ion batteries owing to its high theoretical specific capacity, modest voltage output, and eco-friendly character. It is a challenging task to engineer high-performance composite materials by effectively dispersing Fe3O4 crystals with limited sizes in a well-designed supporting framework following sustainable approaches. In this work, the naturally abundant plant products sodium lignosulfonate (Lig) and sodium cellulose (CMC) were selected to coprecipitate with Fe3+ ions under mild hydrothermal conditions. The Fe-Lig/CMC intermediate sediment with an optimized microstructure can be directly converted to the Lig/CMC-derived carbon matrix-supported Fe3O4 nanocomposite sample (Fe3O4@LigC/CC). Compared with the controlled Fe3O4@LigC material, the Fe3O4@LigC/CC nanocomposite provides superior electrochemical performance in the anode, which has inspiring specific capacities of 820.6 mAh g−1 after 100 cycles under a current rate of 100 mA·g−1 and 750.5 mAh g−1 after 250 cycles, as well as more exciting rate capabilities. The biomimetic sample design and synthesis protocol closely follow the criteria of green chemistry and can be further developed in wider scenarios. Full article
Show Figures

Figure 1

Figure 1
<p>XRD pattern (<b>a</b>), Raman spectrum (<b>b</b>), TGA curve (<b>c</b>), survey XPS spectrum (<b>d</b>), high-resolution C 1s (<b>e</b>), O 1s (<b>f</b>), and Fe 2p (<b>g</b>), XPS spectra of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC sample.</p>
Full article ">Figure 2
<p>FESEM images of the Fe-Lig/CMC (<b>a</b>,<b>b</b>) and Fe<sub>3</sub>O<sub>4</sub>@LigC/CC (<b>c</b>,<b>d</b>) samples and the controlled Fe<sub>3</sub>O<sub>4</sub>@LigC sample (<b>e</b>,<b>f</b>) under corresponding magnifications.</p>
Full article ">Figure 3
<p>TEM (<b>a</b>) and HRTEM (<b>b</b>) images and the EDS testing result (<b>c</b>) of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC sample; illustration of the sustainable and scalable sample design from plant-derived products (<b>d</b>).</p>
Full article ">Figure 4
<p>The CV curves for the initial five cycles (<b>a</b>) and charge and discharge voltage profile (<b>b</b>) of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC sample; comparisons of the low-current cycling performances (<b>c</b>), rate capabilities (<b>d</b>), and high-current cycling performances (<b>e</b>) of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC and Fe<sub>3</sub>O<sub>4</sub>@LigC samples; charge/discharge voltage profile (<b>f</b>) and cycling performance (<b>g</b>) of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC sample in Fe<sub>3</sub>O<sub>4</sub>@LigC/CC//LiFePO<sub>4</sub> full cell.</p>
Full article ">Figure 5
<p>FESEM images of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC electrode before cycling (<b>a</b>,<b>b</b>) and after 250 cycles (<b>c</b>,<b>d</b>) under the corresponding magnifications.</p>
Full article ">Figure 6
<p>Nyquist plot of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC electrode before cycling and after cycling with inset equivalent circuit obtained by fitting the EIS data (<b>a</b>); CV curves of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC electrode at various scan rates from 0.1 to 1.0 mV s<sup>−1</sup> (<b>b</b>); the ln (<span class="html-italic">i</span>)-ln (<span class="html-italic">v</span>) plot (<b>c</b>); capacitive contribution and diffusion ratios at different rates (<b>d</b>); GITT curves of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC electrode (discharge/charge state) (<b>e</b>); voltage (V vs Li<sup>+</sup>/Li) versus time curve for one single GITT test (<b>f</b>).</p>
Full article ">Scheme 1
<p>Schematic diagram of the engineering of the Fe<sub>3</sub>O<sub>4</sub>@LigC/CC nanocomposite.</p>
Full article ">
13 pages, 10800 KiB  
Article
On the Stability of the Interface between Li2TiS3 Cathode and Li6PS5Cl Solid State Electrolytes for Battery Applications: A DFT Study
by Riccardo Rocca, Naiara Leticia Marana, Fabrizio Silveri, Maddalena D’Amore, Eleonora Ascrizzi, Mauro Francesco Sgroi, Nello Li Pira and Anna Maria Ferrari
Batteries 2024, 10(10), 351; https://doi.org/10.3390/batteries10100351 - 7 Oct 2024
Viewed by 383
Abstract
Lithium-titanium-sulfur cathodes have garnered interest due to their distinctive properties and potential applications in lithium-ion batteries. They present various benefits, including lower cost, enhanced safety, and greater energy density compared to the commonly used transition metal oxides. The current trend in lithium-ion batteries [...] Read more.
Lithium-titanium-sulfur cathodes have garnered interest due to their distinctive properties and potential applications in lithium-ion batteries. They present various benefits, including lower cost, enhanced safety, and greater energy density compared to the commonly used transition metal oxides. The current trend in lithium-ion batteries is to move to all-solid-state chemistries in order to improve safety and energy density. Several chemistries for solid electrolytes have been studied, tested, and characterized to evaluate the applicability in energy storage system. Among those, sulfur-based Argyrodites have been coupled with cubic rock-salt type Li2TiS3 electrodes. In this work, Li2TiS3 surfaces were investigated with DFT methods in different conditions, covering the possible configurations that can occur during the cathode usage: pristine, delithiated, and overlithiated. Interfaces were built by coupling selected Li2TiS3 surfaces with the most stable Argyrodite surface, as derived from a previous study, allowing us to understand the (electro)chemical compatibility between these two sulfur-based materials. Full article
Show Figures

Figure 1

Figure 1
<p>(001) Argyrodite surface. The yellow, gray, green, and magenta spheres represent the sulfur, lithium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 2
<p>Schematic representation of LTS surfaces (<b>a</b>) (100), (<b>b</b>) (110), (<b>c</b>) (111). Yellow, gray, and light blue spheres refer to sulfur, lithium, and titanium atoms, respectively.</p>
Full article ">Figure 3
<p>Comparison between the bottom of the conducting band (BCB) and the top of the valence band (TVB) levels for the (<b>a</b>–<b>d</b>) isolated LTS and Argyrodite surfaces, and (<b>e</b>–<b>i</b>) formed interfaces for different LTS lithium content and the two Argyrodite terminations. The lines in blue, red, cyan, and magenta represent the energetic level of pristine LTS and Argyrodite surfaces, and the same surfaces on the formed interfaces, respectively.</p>
Full article ">Figure 4
<p>Final optimized (100)LTS/(001)Argy interfaces (<b>a</b>) LTS/Argy-Li<sub>2</sub>S, (<b>b</b>) LTS/Argy-LPSC, (<b>c</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S, (<b>d</b>) LTS<sub>even</sub>/Argy-Li<sub>2</sub>S, and (<b>e</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S. The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 5
<p>Charge density difference between interface and the pristine sub-units of (<b>a</b>) LTS/Argy-Li<sub>2</sub>S, (<b>b</b>) LTS/Argy-LPSC, (<b>c</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S, (<b>d</b>) LTS<sub>even</sub>/Argy-Li<sub>2</sub>S, and (<b>e</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S. The isosurface corresponds to 0.001 e<sup>−</sup>/bhor<sup>3</sup> with the charge accumulation (depletion) plotted in red (blue). The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">Figure 6
<p>Spin density for the (<b>a</b>) LTS<sub>odd</sub>/Argy-Li<sub>2</sub>S and (<b>b</b>) LTS<sub>over</sub>/Argy-Li<sub>2</sub>S interface. The isosurface corresponds to 0.001 e<sup>−</sup>/bohr<sup>−3</sup>. The yellow, gray, light blue, green, and magenta spheres represent the sulfur, lithium, titanium, chlorine, and phosphorus atoms.</p>
Full article ">
25 pages, 3993 KiB  
Article
Structural and Dynamical Effects of the CaO/SrO Substitution in Bioactive Glasses
by Margit Fabian, Matthew Krzystyniak, Atul Khanna and Zsolt Kovacs
Molecules 2024, 29(19), 4720; https://doi.org/10.3390/molecules29194720 - 5 Oct 2024
Viewed by 497
Abstract
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, [...] Read more.
Silicate glasses containing silicon, sodium, phosphorous, and calcium have the ability to promote bone regeneration and biodegrade as new tissue is generated. Recently, it has been suggested that adding SrO can benefit tissue growth and silicate glass dissolution. Motivated by these recent developments, the effect of SrO/CaO–CaO/SrO substitution on the local structure and dynamics of Si-Na-P-Ca-O oxide glasses has been studied in this work. Differential thermal analysis has been performed to determine the thermal stability of the glasses after the addition of strontium. The local structure has been studied by neutron diffraction augmented by Reverse Monte Carlo simulation, and the local dynamics by neutron Compton scattering and Raman spectroscopy. Differential thermal analysis has shown that SrO-containing glasses have lower glass transition, melting, and crystallisation temperatures. Moreover, the addition of the Sr2+ ions decreased the thermal stability of the glass structure. The total neutron diffraction augmented by the RMC simulation revealed that Sr played a similar role as Ca in the glass structure when substituted on a molar basis. The bond length and the coordination number distributions of the network modifiers and network formers did not change when SrO (x = 0.125, 0.25) was substituted for CaO (25-x). However, the network connectivity increased in glass with 12.5 mol% CaO due to the increased length of the Si-O-Si interconnected chain. The analysis of Raman spectra revealed that substituting CaO with SrO in the glass structure dramatically enhances the intensity of the high-frequency band of 1110–2000 cm−1. For all glasses under investigation, the changes in the relative intensities of Raman bands and the distributions of the bond lengths and coordination numbers upon the SrO substitution were correlated with the values of the widths of nuclear momentum distributions of Si, Na, P, Ca, O, and Sr. The widths of nuclear momentum distributions were observed to soften compared to the values observed and simulated in their parent metal-oxide crystals. The widths of nuclear momentum distributions, obtained from fitting the experimental data to neutron Compton spectra, were related to the amount of disorder of effective force constants acting on individual atomic species in the glasses. Full article
Show Figures

Figure 1

Figure 1
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 1 Cont.
<p>DTA curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 2 Cont.
<p>DTA and DTG curves of Ca25 (<b>a</b>), Ca12.5 (<b>b</b>), and Ca0 (<b>c</b>) glassy samples. See text for details.</p>
Full article ">Figure 3
<p>Experimental and RMC calculated structure factors of Ca25 (red), Ca 12.5(blue), and Ca0 (green) glass samples. Curves are displaced by 1 unit successively for clarity. See text for details.</p>
Full article ">Figure 4
<p>Partial atomic pair correlations for Si-O (<b>a</b>), P-O (<b>b</b>), Ca-O (<b>c</b>), Sr-O (<b>d</b>), Na-O (<b>e</b>), and O-O (<b>f</b>), in glassy samples (Ca25 (red), Ca12.5 (blue), Ca0 (green)). The peak positions corresponding to key bond lengths are shown. See text for details.</p>
Full article ">Figure 5
<p>Si-O (<b>a</b>), P-O (<b>b</b>), CaO (<b>c</b>), and O-O (<b>d</b>) coordination number distributions in the glassy samples.</p>
Full article ">Figure 6
<p>Effects of replacement of CaO with SrO on the Raman spectra of phosphosilicate glasses. See text for details.</p>
Full article ">Figure 7
<p>Fits of the TOF spectra recorded at VESUVIO for bioactive glasses Ca25, Ca12.5, and Ca0. Recoil peaks of individual atomic species in the glasses have been colour-coded, with the recoil peaks due to the aluminium container marked in blue. See text for details.</p>
Full article ">Figure 8
<p>The disorder-induced softening of the widths of nuclear momentum distributions of individual atomic species present in bioactive glasses Ca0, Ca12.5, and Ca25. The bar charts show the adopted disorder scale: (i) the blue bars designate the Maxwell-Boltzmann distribution width limits for completely disordered gas of non-interacting particles without an underlying potential; (ii) the white bars show the upper distribution width limits calculated from atom-projected vibrational densities of states of parent metal oxides; and (iii) the red bars show the widths obtained from the analysis of the NCS experiments. See text for details.</p>
Full article ">
14 pages, 3952 KiB  
Article
Investigating Layered Topological Magnetic Materials as Efficient Electrocatalysts for the Hydrogen Evolution Reaction under High Current Densities
by Sanju Gupta, Hanna Świątek, Mirosław Sawczak, Tomasz Klimczuk and Robert Bogdanowicz
Catalysts 2024, 14(10), 676; https://doi.org/10.3390/catal14100676 - 1 Oct 2024
Viewed by 494
Abstract
Despite considerable progress, high-performing durable catalysts operating under large current densities (i.e., >1000 mA/cm2) are still lacking. To discover platinum group metal-free (PGM-free) electrocatalysts for sustainable energy, our research involves investigating layered topological magnetic materials (semiconducting ferromagnets) as highly efficient electrocatalysts [...] Read more.
Despite considerable progress, high-performing durable catalysts operating under large current densities (i.e., >1000 mA/cm2) are still lacking. To discover platinum group metal-free (PGM-free) electrocatalysts for sustainable energy, our research involves investigating layered topological magnetic materials (semiconducting ferromagnets) as highly efficient electrocatalysts for the hydrogen evolution reaction under high current densities and establishes the novel relations between structure and electrochemical property mechanisms. The materials of interest include transition metal trihalides, i.e., CrCl3, VCl3, and VI3, wherein a structural unit, the layered structure, is formed by Cr (or V) atoms sandwiched between two halides (Cl or I), forming a tri-layer. A few layers of quantum crystals were exfoliated (~50−60 nm), encapsulated with graphene, and electrocatalytic HER tests were conducted in acid (0.5M H2SO4) and alkaline (1M KOH) electrolytes. We find a reasonable HER activity evolved requiring overpotentials in a range of 30–50 mV under 10 mA cm−2 and 400−510 mV (0.5M H2SO4) and 280−500 mV (1M KOH) under −1000 mA cm−2. Likewise, the Tafel slopes range from 27 to 36 mV dec−1 (Volmer–Tafel) and 110 to 190 mV dec−1 (Volmer–Herovsky), implying that these mechanisms work at low and high current densities, respectively. Weak interlayer coupling, spontaneous surface oxidation, the presence of a semi-oxide subsurface (e.g., O–CrCl3), intrinsic Cl (or I) vacancy defects giving rise to in-gap states, electron redistribution (orbital hybridization) affecting the covalency, and sufficiently conductive support interaction lowering the charge transfer resistance endow the optimized adsorption/desorption strength of H* on active sites and favorable electrocatalytic properties. Such behavior is expedited for bi-/tri-layers while exemplifying the critical role of quantum nature electrocatalysts with defect sites for industrial-relevant conditions. Full article
(This article belongs to the Section Catalysis for Sustainable Energy)
Show Figures

Figure 1

Figure 1
<p><b>Morphology and structure.</b> (<b>a</b>–<b>d</b>) Optical photographs; (<b>e</b>–<b>h</b>) scanning electron micrographs; (<b>i</b>) θ–2θ XRD patterns; (<b>j</b>) Micro-Raman spectra excited with 514 nm of few-layered CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub> crystals; and (<b>k</b>) room-temperature electrical conductivity comparison with other reported HER catalysts. Also provided are conventional unit cells of MX<sub>3</sub> crystals with their respective colored atoms, Rietveld refinement, JCPDS nos., and an SEM image of substrate CNW/p–Si (001).</p>
Full article ">Figure 2
<p><b>Cyclic voltammograms and HER polarization curves.</b> (<b>a</b>,<b>b</b>) CV profiles, (<b>c</b>) capacitive current density difference versus scan rate, and (<b>d</b>,<b>e</b>) LSV polarization curves plotted in a large overpotential range with (dotted) and without (solid) considering <span class="html-italic">iR</span> ohmic drop of few-layer CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub> crystals, in acidic (0.5M H<sub>2</sub>SO<sub>4</sub>) and alkaline (1M KOH) electrolytes.</p>
Full article ">Figure 3
<p><b>Overpotential and Tafel slopes under low and high current densities.</b> Overpotential versus log current density, |j|, and Tafel slopes under low current densities of few-layer CrCl<sub>3</sub>, VCl<sub>3</sub>, VI<sub>3</sub>, and VI<sub>2</sub>, in (<b>a</b>) acidic (0.5M H<sub>2</sub>SO<sub>4</sub>) and (<b>b</b>) alkaline (1M KOH) electrolytes. (<b>c</b>) Corresponding Tafel slope analysis under high current densities. The dotted curve shows the coverage-dependent current densities.</p>
Full article ">Figure 4
<p><b>Performance evaluation.</b> Comparison of the required overpotentials to reach the current density of 200 mA cm<sup>−2</sup> and 1000 mA cm<sup>−2</sup> between the recently reported catalysts to those studied here. The scale bar of the catalysts is based on the LSV measurements.</p>
Full article ">Figure 5
<p><b>Proposed HER mechanism in relation to the defects and electronic structure.</b> Illustrations of (<b>a</b>) the experimental scheme for high efficiency HER (<b>b</b>,<b>c</b>), the two-dimensional view of the crystal structure for MX<sub>3</sub> viewed along the (<b>b</b>) <span class="html-italic">a</span>-axis (planar) and (<b>c</b>) <span class="html-italic">c</span>-axis (side view), where the M (=Cr, V) atoms are the bigger spheres, and the X (Cl, I) atoms are the smaller spheres. The MX<sub>6</sub> octahedra form a layered honeycomb lattice via edge-sharing within each layer, and the layers are stacked in an ABC sequence along the <span class="html-italic">c</span>-axis. The bond lengths for CrCl<sub>3</sub> are shown in panel (<b>b</b>). Also provided are the presence of the Cl vacancy (dotted circle) and the vacant interstitial sites occupied by oxygen (solid red circle) in the honeycomb array responsible for the HER reaction. (<b>d</b>) Cartoons of a momentum space diagram and the DOS (density of states) for topological magnetic materials with the coexistence of the WSM and DNL behaviors of the surface/edge states.</p>
Full article ">
15 pages, 5064 KiB  
Article
Adaptation of a Differential Scanning Calorimeter for Simultaneous Electromagnetic Measurements
by John W. Wilson, Mohsen A. Jolfaei, Adam D. Fletcher, Carl Slater, Claire Davis and Anthony J. Peyton
Sensors 2024, 24(18), 6077; https://doi.org/10.3390/s24186077 - 20 Sep 2024
Viewed by 484
Abstract
Although much information can be gained about thermally induced microstructural changes in metals through the measurement of their thermophysical properties using a differential scanning calorimeter (DSC), due to competing influences on the signal, not all microstructural changes can be fully characterised this way. [...] Read more.
Although much information can be gained about thermally induced microstructural changes in metals through the measurement of their thermophysical properties using a differential scanning calorimeter (DSC), due to competing influences on the signal, not all microstructural changes can be fully characterised this way. For example, accurate characterisation of recrystallisation, tempering, and changes in retained delta ferrite in alloyed steels becomes complex due to additional signal changes due to the Curie point, oxidation, and the rate (and therefore the magnitude) of transformation. However, these types of microstructural changes have been shown to invoke strong magnetic and electromagnetic (EM) responses; therefore, simultaneous EM measurements can provide additional complementary data which can help to emphasise or deconvolute these complex signals and develop a more complete understanding of certain metallurgical phenomena. This paper discusses how a DSC machine has been modified to incorporate an EM sensor consisting of two copper coils printed onto either side of a ceramic substrate, with one coil acting as a transmitter and the other as a receiver. The coil is interfaced with a custom-built data acquisition system, which provides current to the transmit coil, records signals from the receive coil, and is controlled by a graphical user interface which allows the user to select multiple excitation frequencies. The equipment has a useable frequency range of approximately 1–100 kHz and outputs phase and magnitude readings at a rate of approximately 50 samples per second. Simultaneous DSC-EM measurements were performed on a nickel sample up to a temperature of 600 °C, with the reversable ferromagnetic to paramagnetic transition in the nickel sample invoking a clear EM response. The results show that the combined DSC-EM apparatus has the potential to provide a powerful tool for the analysis of thermally induced microstructural changes in metals, feeding into research on steel production, development of magnetic and conductive materials, and many more areas. Full article
(This article belongs to the Special Issue Advances and Applications of Magnetic Sensors)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Q2000 DSC, (<b>b</b>) furnace and bowl, (<b>c</b>) close-up of bowl.</p>
Full article ">Figure 2
<p>(<b>a</b>) Ceramic parts for sensor assembly, (<b>b</b>) complete sensor assembly, (<b>c</b>) cross-sectional drawing of DSC coil installation.</p>
Full article ">Figure 3
<p>Block diagram of the EM measurement hardware.</p>
Full article ">Figure 4
<p>A standard transmit waveform comprising 29th and 41st system harmonics. (<b>a</b>) Waveform generated by the system. (<b>b</b>) Transmit current measured by the system current sense ADC. (<b>c</b>) Voltage waveform output to the coil recorded on an external oscilloscope.</p>
Full article ">Figure 5
<p>(<b>a</b>) Setup used to measure the transmit field, (<b>b</b>) close-up of Hall probe and coil showing measurement directions.</p>
Full article ">Figure 6
<p>Large and small test samples.</p>
Full article ">Figure 7
<p>DSC data for Fe-0.5C-6Mn samples with different weights.</p>
Full article ">Figure 8
<p>Test setup for room-temperature laboratory tests.</p>
Full article ">Figure 9
<p>Room-temperature testing. Change in magnitude vs. change in phase for large samples (<b>a</b>) and small samples (<b>b</b>).</p>
Full article ">Figure 10
<p>Room-temperature testing in DSC. Change in magnitude vs. change in phase for large samples (<b>a</b>) and small samples (<b>b</b>), where the solid lines are for tests outside the DSC and dashed lines are for tests inside the DSC.</p>
Full article ">Figure 11
<p>DSC data for a pure nickel sample with a heating and cooling rate of 10 °C/min.</p>
Full article ">Figure 12
<p>EM signal data plotted with respect to the temperature recoded by the DSC for a nickel and no (air) sample. (<b>a</b>) Magnitude at 49 kHz, (<b>b</b>) phase at 49 kHz, (<b>c</b>) magnitude at 71 kHz, (<b>d</b>) phase at 71 kHz.</p>
Full article ">Figure 13
<p>Microstructure of the as-received pure nickel sample before (<b>a</b>) and after (<b>b</b>) the DSC experiment. Sample heated to 600 °C and cooled to 100 °C, with a heating and cooling rate of 10 °C per minute.</p>
Full article ">
10 pages, 2058 KiB  
Article
A WO3–CuCrO2 Tandem Photoelectrochemical Cell for Green Hydrogen Production under Simulated Sunlight
by Ana K. Díaz-García and Roberto Gómez
Molecules 2024, 29(18), 4462; https://doi.org/10.3390/molecules29184462 - 20 Sep 2024
Viewed by 501
Abstract
The development of photoelectrochemical tandem cells for water splitting with electrodes entirely based on metal oxides is hindered by the scarcity of stable p-type oxides and the poor stability of oxides in strongly alkaline and, particularly, strongly acidic electrolytes. As a novelty in [...] Read more.
The development of photoelectrochemical tandem cells for water splitting with electrodes entirely based on metal oxides is hindered by the scarcity of stable p-type oxides and the poor stability of oxides in strongly alkaline and, particularly, strongly acidic electrolytes. As a novelty in the context of transition metal oxide photoelectrochemistry, a bias-free tandem cell driven by simulated sunlight and based on a CuCrO2 photocathode and a WO3 photoanode, both unprotected and free of co-catalysts, is demonstrated to split water while working with strongly acidic electrolytes. Importantly, the Faradaic efficiency for H2 evolution for the CuCrO2 electrode is found to be about 90%, among the highest for oxide photoelectrodes in the absence of co-catalysts. The tandem cell shows no apparent degradation in short-to-medium-term experiments. The prospects of using a practical cell based on this configuration are discussed, with an emphasis on the importance of modifying the materials for enhancing light absorption. Full article
(This article belongs to the Section Electrochemistry)
Show Figures

Figure 1

Figure 1
<p>SEM images corresponding to (<b>A</b>) WO<sub>3</sub> and (<b>B</b>) CuCrO<sub>2</sub> thin-film electrodes.</p>
Full article ">Figure 2
<p>Cyclic voltammograms for (<b>A</b>) WO<sub>3</sub> and (<b>C</b>) CuCrO<sub>2</sub> electrodes in the dark in N<sub>2</sub>-purged 0.1 M HClO<sub>4</sub>. Scan rate 20 mV s<sup>−1</sup>. Linear scan voltammograms for (<b>B</b>) WO<sub>3</sub> and (<b>D</b>) CuCrO<sub>2</sub> electrodes in 0.1 M HClO<sub>4</sub> purged with N<sub>2</sub> under chopped simulated solar illumination (100 mW cm<sup>−2</sup>). Scan rate 5 mV s<sup>−1</sup>.</p>
Full article ">Figure 3
<p>IPCE spectra for (<b>A</b>) CuCrO<sub>2</sub> and (<b>B</b>) WO<sub>3</sub> electrodes in 0.1 M HClO<sub>4</sub> purged with N<sub>2</sub> at 0.10 V<sub>Ag/AgCl</sub> and 0.96 V<sub>Ag/AgCl</sub>, respectively.</p>
Full article ">Figure 4
<p>Linear scan voltammograms under chopped illumination for CuCrO<sub>2</sub> (blue line) and WO<sub>3</sub> (green line) electrodes in 0.1 M HClO<sub>4</sub> purged with N<sub>2</sub>. Note that both cathodic and anodic currents are plotted as positive. The inset shows a detail of the region of curve crossing.</p>
Full article ">Figure 5
<p>(<b>A</b>) Photocurrent vs. voltage curve (recorded at a scan rate of 2 mV s<sup>−1</sup>) for the tandem photoelectrochemical cell in 0.1 M HClO<sub>4</sub> purged with N<sub>2</sub> under illumination with a solar simulator (100 mW cm<sup>−2</sup>). (<b>B</b>) Comparative chronoamperometric curves for the tandem cell in either 0.1 M HClO<sub>4</sub> or 0.5 M Na<sub>2</sub>SO<sub>4</sub> (both purged with N<sub>2</sub>) under simulated solar illumination (100 mW cm<sup>−2</sup>) at zero bias. <span class="html-italic">V</span> = <span class="html-italic">E</span><sub>photoanode</sub> − <span class="html-italic">E</span><sub>photocathode</sub>.</p>
Full article ">Figure 6
<p>Sketch of the different potential levels relevant for the photoelectrodes and the electrolyte for the WO<sub>3</sub>/CuCrO<sub>2</sub> photoelectrochemical tandem device.</p>
Full article ">
34 pages, 21985 KiB  
Review
Emerging Low Detection Limit of Optically Activated Gas Sensors Based on 2D and Hybrid Nanostructures
by Ambali Alade Odebowale, Amer Abdulghani, Andergachew Mekonnen Berhe, Dinelka Somaweera, Sanjida Akter, Salah Abdo, Khalil As’ham, Reza Masoudian Saadabad, Toan T. Tran, David P. Bishop, Alexander S. Solntsev, Andrey E. Miroshnichenko and Haroldo T. Hattori
Nanomaterials 2024, 14(18), 1521; https://doi.org/10.3390/nano14181521 - 19 Sep 2024
Viewed by 1484
Abstract
Gas sensing is essential for detecting and measuring gas concentrations across various environments, with applications in environmental monitoring, industrial safety, and healthcare. The integration of two-dimensional (2D) materials, organic materials, and metal oxides has significantly advanced gas sensor technology, enhancing its sensitivity, selectivity, [...] Read more.
Gas sensing is essential for detecting and measuring gas concentrations across various environments, with applications in environmental monitoring, industrial safety, and healthcare. The integration of two-dimensional (2D) materials, organic materials, and metal oxides has significantly advanced gas sensor technology, enhancing its sensitivity, selectivity, and response times at room temperature. This review examines the progress in optically activated gas sensors, with emphasis on 2D materials, metal oxides, and organic materials, due to limited studies on their use in optically activated gas sensors, in contrast to other traditional gas-sensing technologies. We detail the unique properties of these materials and their impact on improving the figures of merit (FoMs) of gas sensors. Transition metal dichalcogenides (TMDCs), with their high surface-to-volume ratio and tunable band gap, show exceptional performance in gas detection, especially when activated by UV light. Graphene-based sensors also demonstrate high sensitivity and low detection limits, making them suitable for various applications. Although organic materials and hybrid structures, such as metal–organic frameworks (MoFs) and conducting polymers, face challenges related to stability and sensitivity at room temperature, they hold potential for future advancements. Optically activated gas sensors incorporating metal oxides benefit from photoactive nanomaterials and UV irradiation, further enhancing their performance. This review highlights the potential of the advanced materials in developing the next generation of gas sensors, addressing current research gaps and paving the way for future innovations. Full article
(This article belongs to the Special Issue Nanoscale Material-Based Gas Sensors)
Show Figures

Figure 1

Figure 1
<p>Number of publications from 2020 to 2024 based on Scopus: (<b>a</b>) comparison of gas sensor and optically activated gas sensor; (<b>b</b>) different 2D-based gas sensors and metal oxides (internet search of the WoS on 3 June 2024).</p>
Full article ">Figure 2
<p>Schematic illustrating a general structure of an optically activated sensor where the light-assisted reaction induces a measurable change that can detect a target gas, usually in the form of a resistance change that can be further processed in an integrated setup, to detect and monitor the gas in applications of interest.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic diagram of the on-chip graphene photodetector; (<b>b</b>) the graph of minimum detectable concentration (<math display="inline"><semantics> <msub> <mi>c</mi> <mrow> <mi>m</mi> <mi>i</mi> <mi>n</mi> </mrow> </msub> </semantics></math>) versus source power; (<b>c</b>) schematic diagram of overall synthesis processes; (<b>d</b>) the response–recovery curve of GQD gas sensor at 260 nm; (<b>e</b>) the response–recovery curve of GQD gas sensor at 310 nm, with the inset showing the absorption spectra when exposed to air and <math display="inline"><semantics> <msub> <mi>CO</mi> <mn>2</mn> </msub> </semantics></math>; (<b>f</b>) schematic representation of the experimental setup of multilayer graphene (MLG)-based gas sensors; (<b>g</b>) real-time responsiveness towards 1 ppm of <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> of MLG890 under UV@100 (<b>h</b>) MLG935 under UV@100; (<b>i</b>) a comparison of the effect of UV irradiation for MLG890; (<b>j</b>) a comparison of the effect of UV irradiation for MLG935; (<b>k</b>) schematic of fabricated sensor using <math display="inline"><semantics> <msub> <mi>MoS</mi> <mn>2</mn> </msub> </semantics></math>/graphene hybrid; (<b>l</b>) sensor response to 10 ppm <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> concentrations at different working temperatures; (<b>m</b>) sensor response for various <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> concentrations at <math display="inline"><semantics> <mrow> <mn>200</mn> </mrow> </semantics></math> °C; (<b>n</b>) estimated response/recovery time of sensor at <math display="inline"><semantics> <mrow> <mn>200</mn> </mrow> </semantics></math> °C. Reproduced from Refs. [<a href="#B23-nanomaterials-14-01521" class="html-bibr">23</a>,<a href="#B33-nanomaterials-14-01521" class="html-bibr">33</a>,<a href="#B34-nanomaterials-14-01521" class="html-bibr">34</a>,<a href="#B38-nanomaterials-14-01521" class="html-bibr">38</a>].</p>
Full article ">Figure 4
<p>(<b>A</b>) (<b>a</b>) Optical-fiber sensor combined with MOF thin film and gas-sensing system schematic diagram; (<b>b</b>) top and (<b>c</b>) 200 nm ZIF-8-coated optical-fiber cross-sectional FE-SEM pictures showing the continuous and homogeneous thin layer [<a href="#B50-nanomaterials-14-01521" class="html-bibr">50</a>]. (<b>B</b>) (<b>a</b>) Absorbance and (<b>b</b>) variations in the refractive index and thickness of PR-MOF-1 membrane with varying growth periods; (<b>c</b>) hypothesized connection between cut-off thickness and refractive index; and (<b>d</b>) thickness and sensitivity of the PR-MOF-1 membrane [<a href="#B53-nanomaterials-14-01521" class="html-bibr">53</a>].</p>
Full article ">Figure 5
<p>(<b>A</b>) (<b>a</b>) SEM images of PANI/GNF nanocomposite on the SMF; and (<b>b</b>) absorbance spectra of the etched–tapered SMF sensor coated with PANI/GNF nanocomposite towards <math display="inline"><semantics> <msub> <mi>NH</mi> <mn>3</mn> </msub> </semantics></math> in the visible wavelengths range [<a href="#B58-nanomaterials-14-01521" class="html-bibr">58</a>]. (<b>B</b>) (<b>a</b>) Wavelength shift for different types of alcohols; (<b>b</b>) output response of the sensor to the refractive indices of different alcohols [<a href="#B59-nanomaterials-14-01521" class="html-bibr">59</a>].</p>
Full article ">Figure 6
<p>Gas sensors based on MOs heterostructures under UV light activation: (<b>A</b>) (<b>i</b>) hollow nanospheres ZnO/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> heterostructure band diagram; (<b>ii</b>) response of nanospheres ZnO/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> heterojunction sensor to varoius concentrations of <math display="inline"><semantics> <msub> <mi>CH</mi> <mn>2</mn> </msub> </semantics></math>O; (<b>iii</b>) time response of pure ZnO, <math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math>, and ZnO/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> to 5 ppm <math display="inline"><semantics> <msub> <mi>CH</mi> <mn>2</mn> </msub> </semantics></math>O at room temperature under 365 nm LED irradiation (“Reprinted from [<a href="#B101-nanomaterials-14-01521" class="html-bibr">101</a>], Copyright (2023), with permission from Elsevier”). (<b>B</b>) (<b>i</b>) Schematic and energy band diagrams of <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> HNSs sensor in air (top) or TEA gas (bottom); (<b>ii</b>) sensor responses; (<b>iii</b>) response time; and (<b>iv</b>) recovery period of synthesized HNs sensors (“pure <math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> HNs (p-SHNs) and 0.02 mol (0.02 I/SHNs), 0.04 mol (0.04 I/SHNs), and 0.06 mol (0.06 I/SHNs) of <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math> with <math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> HNs”) at different TEA concentrations under 365 nm UV illumination, respectively (“Reprinted from [<a href="#B102-nanomaterials-14-01521" class="html-bibr">102</a>], Copyright (2023), with permission from Elsevier”). (<b>C</b>) (<b>i</b>) Schematic illustration of the <math display="inline"><semantics> <msub> <mi>SO</mi> <mn>2</mn> </msub> </semantics></math> detection mechanism in the 365 nm UV-activated Ag-PANI/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> sensor; (<b>ii</b>) gas-sensing performance of <math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math>, PANI/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math>, and Ag-PANI/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> sensors relative to varying concentrations of <math display="inline"><semantics> <msub> <mi>SO</mi> <mn>2</mn> </msub> </semantics></math> “Reprinted (adapted) with permission from [<a href="#B103-nanomaterials-14-01521" class="html-bibr">103</a>]. Copyright(2021) American Chemical Society”; (<b>iii</b>) schematic illustration of the sensing mechanism in the <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> nanocomposites sensor activated by UV light at room temperature; (<b>iv</b>) response-time curves for <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> nanocomposite sensors to <math display="inline"><semantics> <msub> <mi>CH</mi> <mn>2</mn> </msub> </semantics></math>O concentrations ranging from 0.03 to 10 ppm; and (<b>v</b>) Nyquist plots elucidating the impedance response of the <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> sensor to varying concentrations of <math display="inline"><semantics> <msub> <mi>CH</mi> <mn>2</mn> </msub> </semantics></math>O “Reprinted (adapted) with permission from [<a href="#B104-nanomaterials-14-01521" class="html-bibr">104</a>]. Copyright (2023) American Chemical Society”. (<b>D</b>) Schematic representation of the sensing mechanism in <math display="inline"><semantics> <mrow> <mn>0.01</mn> </mrow> </semantics></math> wt% PANI-loaded <math display="inline"><semantics> <mrow> <mn>0.9</mn> </mrow> </semantics></math>-TiO<sub>2</sub>-<math display="inline"><semantics> <mrow> <mn>0.1</mn> </mrow> </semantics></math>-NiO nanoparticles sensor depicting the different effects of NiO and PANI (<b>i</b>) without UV light, and (<b>ii</b>) with UV light; the selectivity of the sensor at various concentrations of acetone, benzene, toluene, CO, and <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> at (<b>iii</b>) <math display="inline"><semantics> <mrow> <mn>300</mn> </mrow> </semantics></math> °C without UV light and (<b>iv</b>) <math display="inline"><semantics> <mrow> <mn>25</mn> </mrow> </semantics></math> °C with UV light (“Reprinted from [<a href="#B105-nanomaterials-14-01521" class="html-bibr">105</a>], Copyright (2023), with permission from Elsevier”).</p>
Full article ">Figure 7
<p>Gas sensors based on MOs heterostructures under visible-light activation: (<b>A</b>) (<b>i</b>) Illustration of the band diagram and gas-sensing mechanism in ZnO/CsPbBr3; (<b>ii</b>) the repeatability; (<b>iii</b>) the sensor selectivity to various gases (“Reprinted from [<a href="#B108-nanomaterials-14-01521" class="html-bibr">108</a>], Copyright (2022), with permission from Elsevier”). (<b>B</b>) (<b>i</b>) Schematic illustration of band diagram and <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> sensing mechanism under visible-light illumination in <math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math>/S-<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> nanoparticles; (<b>ii</b>) response toward 200–1000 ppb of <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> with L-cysteine amount of 0.6 g (SS2); and (<b>iii</b>) selectivity of SS2 and SS0 (L-cysteine amount of 0 g) under blue-light illumination for various gases [<a href="#B111-nanomaterials-14-01521" class="html-bibr">111</a>]. (<b>C</b>) (<b>i</b>) Enhanced sensing mechanism of <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/ANS/rGO nanohybrids under visible-light illumination; the illustration shows the reaction of HCHO molecules on the surface of the nanohybrids, with the conversion of <math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>2</mn> </msub> </semantics></math> and HCHO into <math display="inline"><semantics> <msub> <mi>CO</mi> <mn>2</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi mathvariant="normal">H</mi> <mn>2</mn> </msub> </semantics></math>O facilitated by changes in the electrical double layer (EDL) thickness and a potential barrier at the <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math> homojunction, which occurs both in air and in the presence of HCHO; energy band diagram of InAG heterojunction in HCHO atmosphere (inset right) and response profiles of the InAG sensor to varying concentrations of HCHO at room temperature (inset left) “Reprinted (adapted)with permission from [<a href="#B112-nanomaterials-14-01521" class="html-bibr">112</a>]. Copyright (2023) American Chemical Society”; (<b>ii</b>) fitted responses of the sensors in [<a href="#B113-nanomaterials-14-01521" class="html-bibr">113</a>] as a function of <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> concentration from 0.02 to 10 ppm at room temperature with visible-light activation; and (<b>iii</b>) schematic representation of the enhanced <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> sensing mechanism by <math display="inline"><semantics> <msub> <mi>In</mi> <mn>2</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">O</mi> <mn>3</mn> </msub> </semantics></math>/g-<math display="inline"><semantics> <msub> <mi mathvariant="normal">C</mi> <mn>3</mn> </msub> </semantics></math><math display="inline"><semantics> <msub> <mi mathvariant="normal">N</mi> <mn>4</mn> </msub> </semantics></math>/Au NFs under visible light (&gt;400 nm and &gt;500 nm) and the corresponding energy band diagrams (“Reprinted from [<a href="#B113-nanomaterials-14-01521" class="html-bibr">113</a>], Copyright (2022), with permission from Elsevier”). (<b>D</b>) Schematic depiction of the sensing mechanisms for (<b>i</b>) N719-<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> films and (<b>ii</b>) N719-PW12/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> films in [<a href="#B115-nanomaterials-14-01521" class="html-bibr">115</a>] under light illumination. The blue, green, and orange spheres denote <math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math>, N719, and PW12 molecules, respectively; (<b>iii</b>) comparison of responses for the sensors to 1 ppm <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> under different light wavelengths; (<b>iv</b>) response and recovery curves of the sensors to 1 ppm <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math>; (<b>v</b>) dynamic response for the PW12/<math display="inline"><semantics> <msub> <mi>TiO</mi> <mn>2</mn> </msub> </semantics></math> sensor under 480 nm light illumination (“Reprinted from [<a href="#B115-nanomaterials-14-01521" class="html-bibr">115</a>], Copyright (2023), with permission from Elsevier”).</p>
Full article ">Figure 8
<p>Gas sensors based on TMDC 2D nanomaterials driven by optical sources: (<b>a</b>) Schematic illustration of sensing mechanism of monolayer (1 L) <math display="inline"><semantics> <msub> <mi>WSe</mi> <mn>2</mn> </msub> </semantics></math> sensor to <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math>; (<b>b</b>) the performance of 1 L <math display="inline"><semantics> <msub> <mi>WSe</mi> <mn>2</mn> </msub> </semantics></math> gas sensor for <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> at RT; and (<b>c</b>) its response fitting curves of different <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> concentrations without and with UV-light irradiations; under UV illumination, the sensor’s response to 1 × <math display="inline"><semantics> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>6</mn> </mrow> </msup> </semantics></math><math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> at room temperature (25 °C) reaches up to 9, marking a fourfold enhancement compared to the sensor’s response without UV activation [<a href="#B121-nanomaterials-14-01521" class="html-bibr">121</a>]; (<b>d</b>) schematic illustration of <math display="inline"><semantics> <msub> <mi>MoS</mi> <mn>2</mn> </msub> </semantics></math>-based sensor, and its relative response under UV light [<a href="#B122-nanomaterials-14-01521" class="html-bibr">122</a>]; schematic sensing process of <math display="inline"><semantics> <msub> <mi>WSe</mi> <mn>2</mn> </msub> </semantics></math> nanosheets in response to <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> (<b>e</b>) without and (<b>f</b>) with UV illumination; (<b>g</b>) the sensing curve to <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> of different concentrations without and with UV illumination; (<b>h</b>) the sensor response of <math display="inline"><semantics> <msub> <mi>WSe</mi> <mn>2</mn> </msub> </semantics></math> nanosheets to <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> with increasing concentration without and with UV illumination; the inset is the linear fit of the sensor response to low <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> concentration [<a href="#B123-nanomaterials-14-01521" class="html-bibr">123</a>]; (<b>i</b>) schematic sensing mechanisms of <math display="inline"><semantics> <msub> <mi>WS</mi> <mn>2</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>WS</mi> <mn>2</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> heterojunction without and with UV illumination; (<b>j</b>) dynamic response curves; and (<b>k</b>) responses and linear fitting curve of <math display="inline"><semantics> <msub> <mi>WS</mi> <mn>2</mn> </msub> </semantics></math>/<math display="inline"><semantics> <msub> <mi>SnO</mi> <mn>2</mn> </msub> </semantics></math> sensor to 0.5–20 ppm <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> at RT under UV illumination [<a href="#B124-nanomaterials-14-01521" class="html-bibr">124</a>]; (<b>l</b>) the <math display="inline"><semantics> <msub> <mi>MoS</mi> <mn>2</mn> </msub> </semantics></math>/ZnO-based actual device attached and wire-bonded with a printed circuit board (PCB); and (<b>m</b>) the response plotted against the concentration of <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> between 5 and 500 ppb and a fitted curve based on the Langmuir adsorption isotherm [<a href="#B125-nanomaterials-14-01521" class="html-bibr">125</a>]; (<b>n</b>) schematic diagram of <math display="inline"><semantics> <msub> <mi>MoTe</mi> <mn>2</mn> </msub> </semantics></math>-based gas sensor; and (<b>o</b>) its dynamic sensing performance with and without UV illumination. The inset shows the dynamic sensing behaviors to <math display="inline"><semantics> <msub> <mi>NO</mi> <mn>2</mn> </msub> </semantics></math> at concentration ranging from 20 to 200 ppb. The light-blue bars represent the concentration of gas for each instance of exposure [<a href="#B126-nanomaterials-14-01521" class="html-bibr">126</a>].</p>
Full article ">
33 pages, 2403 KiB  
Review
Hydrothermal Carbonization of Biomass for Electrochemical Energy Storage: Parameters, Mechanisms, Electrochemical Performance, and the Incorporation of Transition Metal Dichalcogenide Nanoparticles
by Manuel Prieto, Hangbo Yue, Nicolas Brun, Gary J. Ellis, Mohammed Naffakh and Peter S. Shuttleworth
Polymers 2024, 16(18), 2633; https://doi.org/10.3390/polym16182633 - 18 Sep 2024
Viewed by 1065
Abstract
Given the pressing climate and sustainability challenges, shifting industrial processes towards environmentally friendly practices is imperative. Among various strategies, the generation of green, flexible materials combined with efficient reutilization of biomass stands out. This review provides a comprehensive analysis of the hydrothermal carbonization [...] Read more.
Given the pressing climate and sustainability challenges, shifting industrial processes towards environmentally friendly practices is imperative. Among various strategies, the generation of green, flexible materials combined with efficient reutilization of biomass stands out. This review provides a comprehensive analysis of the hydrothermal carbonization (HTC) process as a sustainable approach for developing carbonaceous materials from biomass. Key parameters influencing hydrochar preparation are examined, along with the mechanisms governing hydrochar formation and pore development. Then, this review explores the application of hydrochars in supercapacitors, offering a novel comparative analysis of the electrochemical performance of various biomass-based electrodes, considering parameters such as capacitance, stability, and textural properties. Biomass-based hydrochars emerge as a promising alternative to traditional carbonaceous materials, with potential for further enhancement through the incorporation of extrinsic nanoparticles like graphene, carbon nanotubes, nanodiamonds and metal oxides. Of particular interest is the relatively unexplored use of transition metal dichalcogenides (TMDCs), with preliminary findings demonstrating highly competitive capacitances of up to 360 F/g when combined with hydrochars. This exceptional electrochemical performance, coupled with unique material properties, positions these biomass-based hydrochars interesting candidates to advance the energy industry towards a greener and more sustainable future. Full article
(This article belongs to the Special Issue Carbonized Polymers and Their Functional Applications)
Show Figures

Figure 1

Figure 1
<p>Temperature variation scenarios for the next 100 years [<a href="#B7-polymers-16-02633" class="html-bibr">7</a>], where RCP is ‘Representative Concentration Pathway’. (<b>a</b>) represents CO<sub>2</sub> concentration scenarios and (<b>b</b>) represents global temperature increase scenarios. The dashed line indicates pre-industrial CO<sub>2</sub> concentration.</p>
Full article ">Figure 2
<p>Schematic on hydrothermal carbonization precursors, products, and applications.</p>
Full article ">Figure 3
<p>Mechanistic routes of cellulose and hemicellulose breakdown under hydrothermal conditions. Scheme constructed from information in references [<a href="#B76-polymers-16-02633" class="html-bibr">76</a>,<a href="#B77-polymers-16-02633" class="html-bibr">77</a>,<a href="#B80-polymers-16-02633" class="html-bibr">80</a>].</p>
Full article ">Figure 4
<p>Lignin hydrothermal mechanism. Adapted from [<a href="#B19-polymers-16-02633" class="html-bibr">19</a>] with permission from Elsevier.</p>
Full article ">Figure 5
<p>Energy storage mechanisms in a supercapacitor (adapted from [<a href="#B159-polymers-16-02633" class="html-bibr">159</a>]).</p>
Full article ">
31 pages, 3833 KiB  
Article
Transition Metal-Promoted LDH-Derived CoCeMgAlO Mixed Oxides as Active Catalysts for Methane Total Oxidation
by Marius C. Stoian, Cosmin Romanitan, Katja Neubauer, Hanan Atia, Constantin Cătălin Negrilă, Ionel Popescu and Ioan-Cezar Marcu
Catalysts 2024, 14(9), 625; https://doi.org/10.3390/catal14090625 - 17 Sep 2024
Viewed by 677
Abstract
A series of M(x)CoCeMgAlO mixed oxides with different transition metals (M = Cu, Fe, Mn, and Ni) with an M content x = 3 at. %, and another series of Fe(x)CoCeMgAlO mixed oxides with Fe contents x ranging from 1 to 9 at. [...] Read more.
A series of M(x)CoCeMgAlO mixed oxides with different transition metals (M = Cu, Fe, Mn, and Ni) with an M content x = 3 at. %, and another series of Fe(x)CoCeMgAlO mixed oxides with Fe contents x ranging from 1 to 9 at. % with respect to cations, while keeping constant in both cases 40 at. % Co, 10 at. % Ce and Mg/Al atomic ratio of 3 were prepared via thermal decomposition at 750 °C in air of their corresponding layered double hydroxide (LDH) precursors obtained by coprecipitation. They were tested in a fixed bed reactor for complete methane oxidation with a gas feed of 1 vol.% methane in air to evaluate their catalytic performance. The physico-structural properties of the mixed oxide samples were investigated with several techniques, such as powder X-ray diffraction (XRD), scanning electron microscopy (SEM) coupled with energy dispersive X-ray spectroscopy (EDX), elemental mappings, inductively coupled plasma optical emission spectroscopy (ICP-OES), X-ray photoelectron spectroscopy (XPS), temperature-programmed reduction under hydrogen (H2-TPR) and nitrogen adsorption–desorption at −196 °C. XRD analysis revealed in all the samples the presence of Co3O4 crystallites together with periclase-like and CeO2 phases, with no separate M-based oxide phase. All the cations were distributed homogeneously, as suggested by EDX measurements and elemental mappings of the samples. The metal contents, determined by EDX and ICP-OES, were in accordance with the theoretical values set for the catalysts’ preparation. The redox properties studied by H2-TPR, along with the surface composition determined by XPS, provided information to elucidate the catalytic combustion properties of the studied mixed oxide materials. The methane combustion tests showed that all the M-promoted CoCeMgAlO mixed oxides were more active than the M-free counterpart, the highest promoting effect being observed for Fe as the doping transition metal. The Fe(x)CoCeMgAlO mixed oxide sample, with x = 3 at. % Fe displayed the highest catalytic activity for methane combustion with a temperature corresponding to 50% methane conversion, T50, of 489 °C, which is ca. 40 °C lower than that of the unpromoted catalyst. This was attributed to its superior redox properties and lowest activation energy among the studied catalysts, likely due to a Fe–Co–Ce synergistic interaction. In addition, long-term tests of Fe(3)CoCeMgAlO mixed oxide were performed, showing good stability over 60 h on-stream. On the other hand, the addition of water vapors in the feed led to textural and structural changes in the Fe(3)CoCeMgAlO system, affecting its catalytic performance in methane complete oxidation. At the same time, the catalyst showed relatively good recovery of its catalytic activity as soon as the water vapors were removed from the feed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diffractograms of (<b>a</b>) M(3)CoCeMgAl, and (<b>b</b>) Fe(x)CoCeMgAl LDH-based precursors compared to that of undoped CoCeMgAl LDH. Symbols: #—LDH phase; ∗—boehmite (AlOOH) phase.</p>
Full article ">Figure 2
<p>Diffractograms of (<b>a</b>) M(3)CoCeMgAlO and (<b>b</b>) Fe(x)CoCeMgAlO mixed oxides calcined at 750 °C compared to their unpromoted CoCeMgAlO counterpart. Symbols: Δ—Co<sub>3</sub>O<sub>4</sub> phase; ∗—CeO<sub>2</sub> phase; #—Mg(Al)O phase.</p>
Full article ">Figure 3
<p>(<b>a</b>) High-resolution O 1s core level and (<b>b</b>) C 1s core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide samples: CoCeMgAlO (A); Fe(1)CoCeMgAlO (B); Fe(3)CoCeMgAlO (C); Fe(6)CoCeMgAlO (D); Fe(9)CoCeMgAlO (E).</p>
Full article ">Figure 4
<p>(<b>a</b>) High-resolution Co 2p core level and (<b>b</b>) Ce 3d core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide samples: CoCeMgAlO (A); Fe(1)CoCeMgAlO (B); Fe(3)CoCeMgAlO (C); Fe(6)CoCeMgAlO (D); Fe(9)CoCeMgAlO (E).</p>
Full article ">Figure 5
<p>High-resolution Fe 2p core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide catalysts: Fe(1)CoCeMgAlO (A); Fe(3)CoCeMgAlO (B); Fe(6)CoCeMgAlO (C); Fe(9)CoCeMgAlO (D).</p>
Full article ">Figure 6
<p>H<sub>2</sub>-TPR profiles of CoCeMgAlO and promoted M(3)CoCeMgAlO mixed oxides.</p>
Full article ">Figure 7
<p>H<sub>2</sub>-TPR profiles of the promoted Fe(x)CoCeMgAlO catalysts.</p>
Full article ">Figure 8
<p>The light-off curves for the methane combustion reaction over (<b>a</b>) CoCeMgAlO and M(3)CoCeMgAlO and (<b>b</b>) Fe(x)CoCeMgAlO catalysts. Reaction conditions: 1 vol.% methane in air, GHSV of 16,000 h<sup>−1</sup>, 1 cm<sup>3</sup> of catalyst.</p>
Full article ">Figure 9
<p>Variation of the total hydrogen consumption below 750 °C in the H<sub>2</sub>-TPR measurements and of the intrinsic reaction rates at 400 and 450 °C versus Fe content in the Fe(x)CoCeMgAlO series.</p>
Full article ">Figure 10
<p>(<b>a</b>) Dependence of the Ce/Co surface atomic ratio and of the intrinsic reaction rates at 400 and 450 °C on the Fe content in the Fe(x)CoCeMgAlO series. (<b>b</b>) Dependence between the intrinsic reaction rate at 400 °C and the Ce<sup>4+</sup>/Ce surface atomic ratio in the Fe(x)CoCeMgAlO series.</p>
Full article ">Figure 11
<p>The dependence of the methane total oxidation on the gas hourly space velocity (GHSV) at constant 1 vol. % methane concentration in the feed gas for the Fe(3)CoCeMgAlO catalyst.</p>
Full article ">Figure 12
<p>Evolution of methane conversion at 600 °C with time over Fe(3)CoCeMgAlO catalyst. Reaction conditions: 1 vol.% CH<sub>4</sub> in air and GHSV of 16,000 h<sup>−1</sup> with 1 cm<sup>3</sup> of catalyst.</p>
Full article ">Figure 13
<p>Evolution of methane conversion with time on stream during combustion tests at 600 °C for the Fe(3)CoCeMgAlO catalyst in dry/humid conditions runs. Dry reaction conditions: 1 vol.% CH<sub>4</sub> in air and GHSV of 16,000 h<sup>−1</sup>, 1 cm<sup>3</sup> of catalyst. Humid reaction conditions were obtained by adding, with a peristaltic pump, a flow of 0.14 mL min<sup>−1</sup> of deionized liquid water to the dry mixture, corresponding to a water vapor content of around 40 vol. %.</p>
Full article ">Scheme 1
<p>The scheme for methane catalytic oxidation reaction on the active phase of Co<sub>3</sub>O<sub>4</sub> spinel oxide.</p>
Full article ">
16 pages, 6270 KiB  
Article
C/Co3O4/Diatomite Composite for Microwave Absorption
by Yan Liao, Dashuang Wang, Wenrui Zhu, Zhilan Du, Fanbo Gong, Tuo Ping, Jinsong Rao, Yuxin Zhang and Xiaoying Liu
Molecules 2024, 29(18), 4336; https://doi.org/10.3390/molecules29184336 - 12 Sep 2024
Viewed by 480
Abstract
Transition metal oxides have been widely used in microwave-absorbing materials, but how to improve impedance matching is still an urgent problem. Therefore, we introduced urea as a polymer carbon source into a three-dimensional porous structure modified by Co3O4 nanoparticles and [...] Read more.
Transition metal oxides have been widely used in microwave-absorbing materials, but how to improve impedance matching is still an urgent problem. Therefore, we introduced urea as a polymer carbon source into a three-dimensional porous structure modified by Co3O4 nanoparticles and explored the influence of different heat treatment temperatures on the wave absorption properties of the composite. The nanomaterials, when calcined at a temperature of 450 °C, exhibited excellent microwave absorption capabilities. Specifically, at an optimized thickness of 9 mm, they achieved a minimum reflection loss (RLmin) of −97.3 dB, accompanied by an effective absorption bandwidth (EAB) of 9.83 GHz that comprehensively covered both the S and Ku frequency bands. On the other hand, with a thickness of 3 mm, the RLmin was recorded as −17.9 dB, with an EAB of 5.53 GHz. This excellent performance is attributed to the multi-facial polarization and multiple reflections induced by the magnetic loss capability of Co3O4 nanoparticles, the electrical conductivity of C, and the unique three-dimensional structure of diatomite. For the future development of bio-based microwave absorption, this work provides a methodology and strategy. Full article
(This article belongs to the Special Issue Functional Nanomaterials in Green Chemistry, 2nd Edition)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>–<b>c</b>) SEM images of C/Co<sub>3</sub>O<sub>4</sub>/De-300, (<b>d</b>–<b>f</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-450, and (<b>g</b>–<b>i</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-600; (<b>j</b>–<b>l</b>) TEM images of C/Co<sub>3</sub>O<sub>4</sub>/De-600; and (<b>m</b>–<b>p</b>) HAADF-EDS element mapping images of Co, C, Si, and O of C/Co<sub>3</sub>O<sub>4</sub>/De-450.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns and (<b>b</b>) XPS spectra of full survey scan, (<b>c</b>) C 1s spectrum, (<b>d</b>) Co 2p spectrum, (<b>e</b>) O 1s spectrum, and (<b>f</b>) N 1s spectrum of C/Co<sub>3</sub>O<sub>4</sub>/De-450.</p>
Full article ">Figure 3
<p>(<b>a</b>) FTIR spectra, (<b>b</b>) TGA, and (<b>c</b>) magnetic hysteresis loops for C/Co<sub>3</sub>O<sub>4</sub>/De-300, C/Co<sub>3</sub>O<sub>4</sub>/De-450, and C/Co<sub>3</sub>O<sub>4</sub>/De-600.</p>
Full article ">Figure 4
<p>(<b>a</b>) Real part of permittivity, (<b>b</b>) imaginary part of permittivity, and (<b>c</b>) dielectric loss tangent plots of all the samples; (<b>d</b>) real part of permeability, (<b>e</b>) imaginary part of permeability, and (<b>f</b>) dielectric loss tangent plots of all samples.</p>
Full article ">Figure 5
<p>(<b>a</b>) Cole–Cole plots of C/Co<sub>3</sub>O<sub>4</sub>/De-300, (<b>b</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-450, (<b>c</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-600; plots of (<b>d</b>) impedance matching, (<b>e</b>) attenuation constants, and (<b>f</b>) eddy current loss of all the samples.</p>
Full article ">Figure 6
<p>A 3D loss diagram, contour plot of RL with thickness and frequency (S, C, X and Ku bands), and frequency–loss 1D diagram of (<b>a</b>–<b>d</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-300, (<b>e</b>–<b>h</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-450, and (<b>i</b>–<b>l</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-600.</p>
Full article ">Figure 7
<p>Schematic representation of EMW absorption diagram for C/Co<sub>3</sub>O<sub>4</sub>/De-450 composite.</p>
Full article ">Figure 8
<p>The CST simulation results for (<b>a</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-300, (<b>b</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-450, (<b>c</b>) C/Co<sub>3</sub>O<sub>4</sub>/De-600. (<b>d</b>) The simulated RCS curves of the PEC and C/Co<sub>3</sub>O<sub>4</sub>/De at a scattering angle of 0–180°. (<b>e</b>) The RCS reduction values (the RCS values of PEC minus that of the samples) for all the samples.</p>
Full article ">Figure 9
<p>The synthesis process of C/Co<sub>3</sub>O<sub>4</sub>/De.</p>
Full article ">
Back to TopTop