[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (511)

Search Parameters:
Keywords = nanostructured catalysts

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
12 pages, 2312 KiB  
Article
Aqueous Synthesis of Au10Pt1 Nanorods Decorated with MnO2 Nanosheets for the Enhanced Electrocatalytic Oxidation of Methanol
by Ting Li, Yidan Liu, Yibin Huang, Zhong Yu and Lei Huang
Molecules 2024, 29(16), 3753; https://doi.org/10.3390/molecules29163753 - 7 Aug 2024
Viewed by 319
Abstract
Developing novel catalysts with high activity and high stability for the methanol oxidation reaction (MOR) is of great importance for the ever-broader applications of methanol fuel cells. Herein, we present a facile technique for synthesizing Au10Pt1@MnO2 catalysts using [...] Read more.
Developing novel catalysts with high activity and high stability for the methanol oxidation reaction (MOR) is of great importance for the ever-broader applications of methanol fuel cells. Herein, we present a facile technique for synthesizing Au10Pt1@MnO2 catalysts using a wet chemical method and investigate their catalytic performance for the MOR. Notably, the Au10Pt1@MnO2-M composite demonstrated a significantly high peak mass activity of 15.52 A mg(Pt)−1, which is 35.3, 57.5, and 21.9 times greater than those of the Pt/C (0.44 A mg(Pt)−1), Pd/C (0.27 A mg(Pt)−1), and Au10Pt1 (0.71 A mg(Pt)−1) catalysts, respectively. Comparative analysis with commercial Pt/C and Pd/C catalysts, as well as Au10Pt1 HSNRs, revealed that the Au10Pt1@MnO2-M composite exhibited the lowest initial potential, the highest peak current density, and superior CO anti-poisoning capability. The results demonstrate that the introduction of MnO2 nanosheets, with excellent oxidation capability, not only significantly increases the reactive sites, but also promotes the reaction kinetics of the catalyst. Furthermore, the high surface area of the MnO2 nanosheets facilitates charge transfer and induces modifications in the electronic structure of the composite. This research provides a straightforward and effective strategy for the design of efficient electrocatalytic nanostructures for MOR applications. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) UV–VIS absorption spectra of the samples; (<b>b</b>) TEM image of Au NPs, (<b>c</b>) Au<sub>10</sub>Pt<sub>1</sub> HSNRs, and (<b>d</b>) Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-M; local amplification of HRTEM images of (<b>e</b>) Au<sub>10</sub>Pt<sub>1</sub> HSNRs and (<b>f</b>) MnO<sub>2</sub> nanosheets; (<b>g</b>) selected area electron diffraction (SAED) image.</p>
Full article ">Figure 2
<p>XPS spectra of Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-M: (<b>a</b>) survey; (<b>b</b>) Au; (<b>c</b>) Pt; (<b>d</b>) Mn.</p>
Full article ">Figure 3
<p>TEM images of different KMnO<sub>4</sub> additions: (<b>a</b>) Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-L; (<b>b</b>) Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-M; (<b>c</b>) Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-H.</p>
Full article ">Figure 4
<p>TEM images of Au<sub>10</sub>Pt<sub>1</sub>@MnO<sub>2</sub>-M at different reaction temperatures: (<b>a</b>) 35 °C; (<b>b</b>) 45 °C; (<b>c</b>) 60 °C.</p>
Full article ">Figure 5
<p>The performance of different catalysts for MOR in 1 M KOH and 1 M CH<sub>3</sub>OH solutions: (<b>a</b>) CV curve of specific activity; (<b>b</b>) CV curve per mass of noble metals; (<b>c</b>) mass activity at the highest current density; (<b>d</b>) chronoamperometry measurement curves.</p>
Full article ">
12 pages, 6701 KiB  
Article
Construction of a Wood Nanofiber–Bismuth Halide Photocatalyst and Catalytic Degradation Performance of Tetracycline from Aqueous Solutions
by Jiarong She, Cuihua Tian, Yan Qing and Yiqiang Wu
Molecules 2024, 29(14), 3253; https://doi.org/10.3390/molecules29143253 - 10 Jul 2024
Viewed by 630
Abstract
Nanostructured bismuth oxide bromide (BiOBr) has attracted considerable attention as a visible light catalyst. However, its photocatalytic degradation efficiency is limited by its low specific surface area. In this study, a solvothermal approach was employed to synthesize BiOBr, which was subsequently loaded onto [...] Read more.
Nanostructured bismuth oxide bromide (BiOBr) has attracted considerable attention as a visible light catalyst. However, its photocatalytic degradation efficiency is limited by its low specific surface area. In this study, a solvothermal approach was employed to synthesize BiOBr, which was subsequently loaded onto cellulose nanofibers (CNFs) to obtain a bismuth halide composite catalyst. The performance of this catalyst in the removal of refractory organic pollutants such as tetracycline (TC) from solutions under visible light excitation was examined. Our results indicate that BiOBr/CNF effectively removes TC from the solution under light conditions. At a catalyst dosage of 100 mg/L, the removal efficiency for TC (with an initial concentration of 100 mg/L) was 94.2%. This study elucidates the relationship between the microstructure of BiOBr/CNF composite catalysts and their improved photocatalytic activity, offering a new method for effectively removing pollutants from water. Full article
(This article belongs to the Special Issue Advanced Materials in Photoelectrochemistry)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) X-ray diffraction patterns and (<b>b</b>) Fourier transform infrared spectra of the prepared samples.</p>
Full article ">Figure 2
<p>Scanning electron microscopy images of (<b>a</b>) cellulose nanofibers (CNF), (<b>b</b>) BiOBr/CNF.</p>
Full article ">Figure 3
<p>Transmission electron microscopy images of (<b>a</b>) CNF and (<b>b</b>) BiOBr/CNF.</p>
Full article ">Figure 4
<p>UV–vis diffused reflectance spectra of samples.</p>
Full article ">Figure 5
<p>The I-T curves (<b>a</b>) and the EIS response (<b>b</b>) of sample.</p>
Full article ">Figure 6
<p>Adsorption and degradation performance of tetracycline in various samples under (<b>a</b>) light avoidance, (<b>b</b>) visible light exposure, and (<b>c</b>) UV light exposure.</p>
Full article ">Figure 7
<p>Effects of (<b>a</b>) pH, (<b>b</b>) initial TC concentration on removal efficiency, and (<b>c</b>) catalyst dosage.</p>
Full article ">Figure 8
<p>Free radical trapping with different free radical scavenger.</p>
Full article ">Figure 9
<p>Photocatalytic degradation mechanism of tetracycline by CNF/BiOBr.</p>
Full article ">Figure 10
<p>Possible degradation pathway of the CNF/BiOBr composite.</p>
Full article ">Figure 11
<p>Schematic diagram of the preparation route of the BiOBr/CNF.</p>
Full article ">
13 pages, 3145 KiB  
Article
Self-Reconstructed Metal–Organic Framework-Based Hybrid Electrocatalysts for Efficient Oxygen Evolution
by Kunting Cai, Weibin Chen, Yinji Wan, Hsingkai Chu, Xiao Hai and Ruqiang Zou
Nanomaterials 2024, 14(14), 1168; https://doi.org/10.3390/nano14141168 - 9 Jul 2024
Viewed by 598
Abstract
Refining synthesis strategies for metal–organic framework (MOF)-based catalysts to improve their performance and stability in an oxygen evolution reaction (OER) is a big challenge. In this study, a series of nanostructured electrocatalysts were synthesized through a solvothermal method by growing MOFs and metal–triazolates [...] Read more.
Refining synthesis strategies for metal–organic framework (MOF)-based catalysts to improve their performance and stability in an oxygen evolution reaction (OER) is a big challenge. In this study, a series of nanostructured electrocatalysts were synthesized through a solvothermal method by growing MOFs and metal–triazolates (METs) on nickel foam (NF) substrates (named MET-M/NF, M = Fe, Co, Cu), and these electrocatalysts could be used directly as OER self-supporting electrodes. Among these electrocatalysts, MET-Fe/NF exhibited the best OER performance, requiring only an overpotential of 122 mV at a current density of 10 mA cm−2 and showing remarkable stability over 15 h. The experimental results uncovered that MET-Fe/NF underwent an in situ structural reconstruction, resulting in the formation of numerous iron/nickel (oxy)hydroxides with high OER activity. Furthermore, in a two-electrode water-splitting setup, MET-Fe/NF only required 1.463 V to achieve a current density of 10 mA cm−2. Highlighting its potential for practical applications. This work provides insight into the design and development of efficient MOF-based OER catalysts. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustrating the synthesis process and in situ structure reconstruction of the MET-Fe/NF catalyst.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD pattern of MET-Fe/NF (inset: a detailed view of the dotted area). (<b>b</b>) N<sub>2</sub> adsorption-desorption isotherm of MET-Fe (inset: pore size distribution curve).</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>c</b>) SEM images of MET-Fe/NF at different magnifications. (<b>d</b>) Elemental mapping images of MET-Fe/NF.</p>
Full article ">Figure 4
<p>Electrocatalytic OER performance. (<b>a</b>) LSV curves. (<b>b</b>) Comparison of the overpotentials at 10 mA cm<sup>−2</sup>. (<b>c</b>) Tafel plots. (<b>d</b>) Electrochemical double-layer capacitance. (<b>e</b>) Stability test at constant current density of 10 mA cm<sup>−2</sup> without iR compensation. (<b>f</b>) Comparison of the OER activity of MET-Fe/NF and other reported catalysts: overpotential at 10 mA cm<sup>−2</sup> and the corresponding Tafel slope.</p>
Full article ">Figure 5
<p>(<b>a</b>) LSV curve and photograph of overall water splitting over the MET-Fe/NF||Pt/C/NF two-electrode setup. (<b>b</b>) Stability test of overall water splitting at constant current density of 10 mA cm<sup>−2</sup> without iR compensation.</p>
Full article ">Figure 6
<p>(<b>a</b>) SEM images of MET-Fe/NF after OER test. (<b>b</b>) XRD pattern of MET-Fe/NF after OER test. High-resolution XPS spectra of (<b>c</b>) Fe 2p and (<b>d</b>) C 1s of MET-Fe/NF before and after OER test.</p>
Full article ">
16 pages, 11294 KiB  
Article
Metal-Catalyzed Thermo-Catalytic Decomposition and Continuous Catalyst Generation
by Mpila Makiesse Nkiawete and Randy Lee Vander Wal
Catalysts 2024, 14(7), 414; https://doi.org/10.3390/catal14070414 - 29 Jun 2024
Viewed by 519
Abstract
In this study, metal dusting is utilized to initiate a two-stage thermo-catalytic decomposition (TCD) process. Stage 1 starts with metal-catalyzed TCD, and in stage 2 the metal-catalyzed carbon catalyzes additional TCD. TEM is presented of the early- versus late-stage TCD to qualitatively illustrate [...] Read more.
In this study, metal dusting is utilized to initiate a two-stage thermo-catalytic decomposition (TCD) process. Stage 1 starts with metal-catalyzed TCD, and in stage 2 the metal-catalyzed carbon catalyzes additional TCD. TEM is presented of the early- versus late-stage TCD to qualitatively illustrate the second-stage TCD by the metal-catalyzed carbons. Corresponding SEM illustrates differences in growth type and surface density between early versus late reaction times, with backscattered imaging differentiating the first- versus second-stage TCD. TGA supports the microscopic inference of a second carbon phase by the presence of an early (low-temperature) reaction peak, characteristic of low-structure or disordered carbon as the second-stage TCD carbon. Raman analysis confirms that the second-stage carbon deposit is more disordered and unstructured, especially at 1000 °C, supported by the ID/IG and La value changes from 0.068 to 0.936 and 65 nm to 4.7 nm, respectively. To further confirm second-stage TCD occurrence upon pre-catalyzed carbons, two carbon blacks are tested. Exposing a combination of edge and basal or exclusively basal sites for the graphitized form, they afford a direct comparison of TCD carbon nanostructure dependence upon the initial carbon catalyst nanostructure. Pre-oxidation of the stainless-steel wool (SSW) prior to TCD is advantageous, accelerating TCD rates and increasing carbon yield relative to the nascent SSW for an equivalent reaction duration. Full article
(This article belongs to the Section Industrial Catalysis)
Show Figures

Figure 1

Figure 1
<p>Representation of the 2-stage catalytic TCD process.</p>
Full article ">Figure 2
<p>(<b>a</b>). A series of SEM images at 5, 10, and 20 min of TCD duration illustrating the collective concurrence of the sequential reaction processes. (<b>b</b>). SEM images of the SS mesh after 1 h of TCD with SNG, revealing the dense combination of catalyzed carbon filament growth and carbon deposition.</p>
Full article ">Figure 3
<p>Metal dusting array of carbon flora, ranging from rosebuds to ridges to fibers to CNTs of varied types. Arrows indicate the different carbon forms that are generated.</p>
Full article ">Figure 4
<p>(<b>a</b>) Spotlighting secondary carbon deposition by contrasting HRTEM images of a CNT at early and later stages of TCD. (<b>b</b>) Spotlighting secondary carbon deposition by contrasting HRTEM images of a graphene stack at an early and later stage of TCD. Arrows indicate the different carbon forms that are generated.</p>
Full article ">Figure 5
<p>TGA oxidation of 1st- and 2nd-stage depositions, illustrating the different carbon phases through the different inset temperatures and derivatives.</p>
Full article ">Figure 6
<p>Raman spectroscopy highlighting the different nanostructures of the TCD carbon (increasingly disordered—2nd stages (<b>b</b>,<b>c</b>)) and initial carbon deposit by metal dusting (ordered/graphitic—1st stage (<b>a</b>)) with inserts illustrating the different carbon phases through TGA weight loss derivatives.</p>
Full article ">Figure 7
<p>SEM backscattered electron imaging showing (<b>a</b>) the presence of metal particles (bright spots) during the first-stage metal dusting reaction at 900 °C and (<b>b</b>) the near absence of observable metal particles during the second-stage deposition at 1000 °C.</p>
Full article ">Figure 8
<p>HRTEM images of nascent R250 and its graphitized analogue (G-R250) contrasting their different nanostructures.</p>
Full article ">Figure 9
<p>HRTEM images of the carbon deposits upon the carbon black catalysts post-TCD. The irregular morphology, resembling sea coral, and disordered lamellae therein appear common to both catalysts, despite their very different nanostructures.</p>
Full article ">Figure 10
<p>SEM images of nascent and rusted SSW post-TCD deposits using a mixture of 25% SNG with balance argon at 900 °C for 30 min.</p>
Full article ">Figure 11
<p>Metal dusting apparatus.</p>
Full article ">
13 pages, 3067 KiB  
Article
CO2 Electroreduction by Engineering the Cu2O/RGO Interphase
by Matteo Bisetto, Sourav Rej, Alberto Naldoni, Tiziano Montini, Manuela Bevilacqua and Paolo Fornasiero
Catalysts 2024, 14(7), 412; https://doi.org/10.3390/catal14070412 - 28 Jun 2024
Viewed by 471
Abstract
In the present investigation, Cu2O-based composites were successfully prepared through a multistep method where cubic Cu2O nanoparticles (CU Cu2O) have been grown on Reduced Graphene Oxide (RGO) nanosheets. The structural and morphological properties of the materials have [...] Read more.
In the present investigation, Cu2O-based composites were successfully prepared through a multistep method where cubic Cu2O nanoparticles (CU Cu2O) have been grown on Reduced Graphene Oxide (RGO) nanosheets. The structural and morphological properties of the materials have been studied through a comprehensive characterization, confirming the coexistence of crystalline Cu2O and RGO. Microscopical imaging revealed the intimate contact between the two materials, affecting the size and the distribution of Cu2O nanoparticles on the support. The features of the improved morphology strongly affected the electrochemical behavior of the composites, increasing the activity and the faradaic efficiencies towards the electrochemical CO2 reduction reaction process. CU Cu2O/RGO 2:1 composite displayed selective CO formation over H2, with higher currents compared to pristine Cu2O (−0.34 mA/cm2 for Cu2O and −0.64 mA/cm2 for CU Cu2O/RGO 2:1 at the voltage of −0.8 vs. RHE and in a CO2 atmosphere) and a faradaic efficiency of 50% at −0.9 V vs. RHE. This composition exhibited significantly higher CO production compared to the pristine materials, indicating a favorable *CO intermediate pathway even at lower voltages. The systematic investigation on the effects of nanostructuration on composition, morphology and catalytic behavior is a valuable solution for the formation of effective interphases for the promotion of catalytic properties providing crucial insights for future catalysts design and applications. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Experimental procedure for the synthesis of cubic Cu<sub>2</sub>O/RGO electrocatalysts.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD patterns for pristine nanoparticles of Cu<sub>2</sub>O and for the composite with RGO. (<b>b</b>) Raman spectra for the different synthetized composites where it is possible to observe the characteristic vibrations of Cu<sub>2</sub>O (#) and RGO (*). (<b>c</b>) Thermal Gravimetric Analysis for the three different composites in air with the temperature range 50–800 °C.</p>
Full article ">Figure 3
<p>Microscope analysis for the composite Cu<sub>2</sub>O/RGO 2:1. (<b>a</b>,<b>b</b>) TEM images at different magnification levels. (<b>c</b>) SEM image of the cubic nanoparticles of Cu<sub>2</sub>O growth on RGO surface. (<b>d</b>) Size distribution of Cu<sub>2</sub>O nanoparticles (count: 400 nanoparticles). (<b>e</b>) EDX image and elemental distribution for the considered composite.</p>
Full article ">Figure 4
<p>CO<sub>2</sub>RR tests for the different composites of CU Cu<sub>2</sub>O/RGO at the voltage of −0.9 V vs. RHE. (<b>a</b>) Two hour chronoamperometries under saturated CO<sub>2</sub> atmosphere. (<b>b</b>) FEs of gaseous and liquid products. (<b>c</b>) Products distribution (μmol) of the different compounds after 2 h chronoamperometries. (<b>d</b>) Production profile of CO during the experiment for the different composite materials.</p>
Full article ">Figure 5
<p>Faradaic efficiencies at the different voltages for the three composite CU Cu<sub>2</sub>O/RGO. (<b>a</b>) CU Cu<sub>2</sub>O/RGO 1:2, (<b>b</b>) CU Cu<sub>2</sub>O/RGO 1:1 and (<b>c</b>) CU Cu<sub>2</sub>O/RGO 2:1.</p>
Full article ">
16 pages, 4672 KiB  
Review
Lattice-Strained Bimetallic Nanocatalysts: Fundamentals of Synthesis and Structure
by Yaowei Wang, Huibing Shi, Deming Zhao, Dongpei Zhang, Wenjuan Yan and Xin Jin
Molecules 2024, 29(13), 3062; https://doi.org/10.3390/molecules29133062 - 27 Jun 2024
Viewed by 410
Abstract
Bimetallic nanostructured catalysts have shown great promise in the areas of energy, environment and magnetics. Tunable composition and electronic configurations due to lattice strain at bimetal interfaces have motivated researchers worldwide to explore them industrial applications. However, to date, the fundamentals of the [...] Read more.
Bimetallic nanostructured catalysts have shown great promise in the areas of energy, environment and magnetics. Tunable composition and electronic configurations due to lattice strain at bimetal interfaces have motivated researchers worldwide to explore them industrial applications. However, to date, the fundamentals of the synthesis of lattice-mismatched bimetallic nanocrystals are still largely uninvestigated for most supported catalyst materials. Therefore, in this work, we have conducted a detailed review of the synthesis and structural characterization of bimetallic nanocatalysts, particularly for renewable energies. In particular, the synthesis of Pt, Au and Pd bimetallic particles in a liquid phase has been critically discussed. The outcome of this review is to provide industrial insights of the rational design of cost-effective nanocatalysts for sustainable conversion technologies. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>Au and Pt composites are the most promising catalysts for the future chemical industry.</p>
Full article ">Figure 2
<p>Schematic representation of some possible mixing patterns in bimetallic systems: (<b>a</b>) core–shell alloys, (<b>b</b>) sub-cluster segregated alloys, (<b>c</b>) ordered and random homogeneous alloys, and (<b>d</b>) multishell alloys. Reproduced with permission from Ref. [<a href="#B21-molecules-29-03062" class="html-bibr">21</a>]. Copyright 2012, Royal Society of Chemistry.</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic of different methods for encapsulating metal within MFI zeolite. Reproduced with permission from Ref. [<a href="#B27-molecules-29-03062" class="html-bibr">27</a>]. Copyright 2018, Wiley. (<b>b</b>) Schematic illustration of the fabrication process of nanoporous Pt–Co alloy nanowires. Reproduced with permission from Ref. [<a href="#B29-molecules-29-03062" class="html-bibr">29</a>]. Copyright 2009, American Chemical Society.</p>
Full article ">Figure 4
<p>(<b>a</b>) UV–VIS spectra of Au, Ag and Ag-Au bimetallic nanocomposites synthesized by the graft copolymer HES-g-poly. Reproduced with permission from Ref. [<a href="#B50-molecules-29-03062" class="html-bibr">50</a>]. Copyright 2017, Elsevier. (<b>b</b>) DRIFT spectra of Au–Ag/TiO<sub>2</sub> after in situ reduction and CO adsorption. Evolution of the spectra with CO contact time from 0 to 80 min (a–h). Reproduced with permission from Ref. [<a href="#B49-molecules-29-03062" class="html-bibr">49</a>]. Copyright 2015, Elsevier.</p>
Full article ">Figure 5
<p>(<b>a</b>) HAADF-STEM image of Au<sub>3</sub>-Cu/rGO catalyst; (<b>b</b>–<b>d</b>) EDS mapping of Au<sub>3</sub>-Cu/rGO catalyst. Reproduced with permission from Ref. [<a href="#B52-molecules-29-03062" class="html-bibr">52</a>] Copyright 2017, Elsevier; (<b>e</b>) typical HRTEM images of Pd/Au alloy NPs in the as-obtained Pd/Au@g-C<sub>3</sub>N<sub>4</sub>-N(1:1). Reproduced with permission from Ref. [<a href="#B45-molecules-29-03062" class="html-bibr">45</a>] Copyright 2017, Elsevier; (<b>f</b>) HRTEM image of Pd@Ag/RGO. Reproduced with permission from Ref. [<a href="#B46-molecules-29-03062" class="html-bibr">46</a>]. Copyright 2018, Elsevier; (<b>g</b>) HRSEM images of AuPt-BNP/FTO glass. Reproduced with permission from Ref. [<a href="#B48-molecules-29-03062" class="html-bibr">48</a>]. Copyright 2015, Elsevier; (<b>h</b>) TEM image of ZnO nanopyramids. Reproduced with permission from Ref. [<a href="#B54-molecules-29-03062" class="html-bibr">54</a>]. Copyright 2015, Elsevier.</p>
Full article ">Figure 6
<p>(<b>a</b>) Emission spectra of monometallic and bimetallic nanocomposites modified TiO<sub>2</sub>. (<b>b</b>) Effect of different amounts of CuSO<sub>4</sub> (0.01 M) deposition onto Au nanospheres for variation in the surface plasmon band, and (<b>c</b>) their respective color changes (increasing Cu<sup>2+</sup> ions from A to D). Reproduced with permission from Ref. [<a href="#B53-molecules-29-03062" class="html-bibr">53</a>]. Copyright 2017, Elsevier.</p>
Full article ">Figure 7
<p>Lattice-mismatched anisotropic growth of PtFe nanoclusters. HR-TEM images of (<b>a</b>) Pt and (<b>b</b>) PtFe(1) with insets at lower magnification. White bars indicate 5 nm. Reproduced with permission from Ref. [<a href="#B56-molecules-29-03062" class="html-bibr">56</a>]. Copyright 2016, Elsevier.</p>
Full article ">Figure 8
<p>TEM images of bimetallic PtMn catalysts. (<b>a</b>–<b>c</b>) novel nano-bud shaped bimetallic clusters, (<b>d</b>,<b>e</b>) 4 to 8 buds on the PtMn clusters. Anisotropic growth orienting from the surface plane of ordered Pt octahedral or cubic structures. (<b>f</b>) Lattice-strain-induced distorted bimetallic PtMn nanocatalysts. Reproduced with permission from Ref. [<a href="#B63-molecules-29-03062" class="html-bibr">63</a>]. Copyright 2017, Elsevier.</p>
Full article ">
18 pages, 6181 KiB  
Article
Growth of Carbon Nanofibers and Carbon Nanotubes by Chemical Vapour Deposition on Half-Heusler Alloys: A Computationally Driven Experimental Investigation
by Ioannis G. Aviziotis, Apostolia Manasi, Afroditi Ntziouni, Georgios P. Gakis, Aikaterini-Flora A. Trompeta, Xiaoying Li, Hanshan Dong and Costas A. Charitidis
Materials 2024, 17(13), 3144; https://doi.org/10.3390/ma17133144 - 27 Jun 2024
Viewed by 464
Abstract
The possibility of directly growing carbon nanofibers (CNFs) and carbon nanotubes (CNTs) on half-Heusler alloys by Chemical Vapour Deposition (CVD) is investigated for the first time, without using additional catalysts, since the half-Heusler alloys per se may function as catalytic substrates, according to [...] Read more.
The possibility of directly growing carbon nanofibers (CNFs) and carbon nanotubes (CNTs) on half-Heusler alloys by Chemical Vapour Deposition (CVD) is investigated for the first time, without using additional catalysts, since the half-Heusler alloys per se may function as catalytic substrates, according to the findings of the current study. As a carbon source, acetylene is used in the temperature range of 700–750 °C. The n-type half-Heusler compound Zr0.4Ti0.60.33Ni0.33Sn0.98Sb0.020.33 is utilized as the catalytic substrate. At first, a computational model is developed for the CVD reactor, aiming to optimize the experimental process design and setup. The experimental process conditions are simulated to investigate the reactive species concentrations within the reactor chamber and the activation of certain reactions. SEM analysis confirms the growth of CNFs with diameters ranging from 450 nm to 1 μm. Raman spectroscopy implies that the formed carbon structures resemble CNFs rather than CNTs, and that amorphous carbon also co-exists in the deposited samples. From the characterization results, it may be concluded that a short reaction time and a low acetylene flow rate lead to the formation of a uniform CNF coating on the surface of half-Heusler alloys. The purpose of depositing carbon nanostructures onto half-Heusler alloys is to improve the current transfer, generated from these thermoelectric compounds, by forming a conductive coating on their surface. Full article
Show Figures

Figure 1

Figure 1
<p>SEM/EDS analysis on the surface of the half-Heulser alloy on three different spots, marked with coloured crosses.</p>
Full article ">Figure 2
<p>The XRD pattern of the <math display="inline"><semantics> <mrow> <msub> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi mathvariant="normal">Z</mi> <mi mathvariant="normal">r</mi> </mrow> <mrow> <mn>0.4</mn> </mrow> </msub> <msub> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">i</mi> </mrow> <mrow> <mn>0.6</mn> </mrow> </msub> </mrow> </mfenced> </mrow> <mrow> <mn>0.33</mn> </mrow> </msub> <msub> <mrow> <mi mathvariant="normal">N</mi> <mi mathvariant="normal">i</mi> </mrow> <mrow> <mn>0.33</mn> </mrow> </msub> <msub> <mrow> <mfenced separators="|"> <mrow> <msub> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">n</mi> </mrow> <mrow> <mn>0.98</mn> </mrow> </msub> <msub> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">b</mi> </mrow> <mrow> <mn>0.02</mn> </mrow> </msub> </mrow> </mfenced> </mrow> <mrow> <mn>0.33</mn> </mrow> </msub> </mrow> </semantics></math> half-Heusler alloy prior to and after sintering.</p>
Full article ">Figure 3
<p>(<b>a</b>) The velocity vector and streamlines and (<b>b</b>) the temperature distribution within the CVD reactor, for the simulated case using the conditions of HH1.</p>
Full article ">Figure 4
<p>Acetylene mole fraction within the reactor for different oven temperatures: (<b>a</b>) 700 °C, (<b>b</b>) 900 °C.</p>
Full article ">Figure 5
<p>Acetylene/byproduct ratio as a function of temperature.</p>
Full article ">Figure 6
<p>Effect of the inflow rate on the acetylene/byproduct ratio. (<b>a</b>) Varying acetylene flow with a constant N<sub>2</sub> flow of 180 sccm and (<b>b</b>) varying N<sub>2</sub> flow with a constant acetylene flow of 46.4 sccm.</p>
Full article ">Figure 7
<p>Indicative photos of the half-Heusler sample HH2 (<b>a</b>) before and (<b>b</b>) after deposition.</p>
Full article ">Figure 8
<p>SEM micrographs for sample HH1 after deposition where carbon deposition can be observed at different areas of HH1 (<b>a</b>), and long tubular structures with thick diameters are observed (<b>b</b>), which is a magnification of (<b>a</b>).</p>
Full article ">Figure 9
<p>Raman spectrum of HH1 where the D, G, and 2D bands are shown at 1340 cm<sup>−1</sup>, 1606 cm<sup>−1</sup>, and 2679 cm<sup>−1</sup>, respectively, and the additional D+G and 2D’ bands at 2918 cm<sup>−1</sup> and 3197 cm<sup>−1</sup>, respectively.</p>
Full article ">Figure 10
<p>Raman spectra of HH1 (red) and HH2 (black) where it is shown that the characteristic bands of the two samples are at the same Raman shift.</p>
Full article ">Figure 11
<p>SEM micrographs of HH3 where (<b>a</b>) absent and (<b>b</b>) limited cylindrical carbon nanostructures are shown.</p>
Full article ">Figure 12
<p>SEM micrographs of HH4 where coating uniformity is shown in (<b>a</b>,<b>b</b>) and CNFs are visible from (<b>c</b>,<b>d</b>) with the diameter size ranging between 800 and 1000 nm (<b>d</b>).</p>
Full article ">Figure 13
<p>Raman spectrum of HH4 where the D, G, and 2D bands are shown at 1346 cm<sup>−1</sup>, 1600cm<sup>−1</sup>, and 2693 cm<sup>−1</sup>, and the additional D+G and 2D’ bands at 2932 cm<sup>−1</sup> and 3217 cm<sup>−1</sup>, respectively.</p>
Full article ">Figure 14
<p>SEM micrographs of HH5 where coating uniformity is shown in (<b>a</b>,<b>b</b>), and potentially formed MWCNTs are visible from (<b>c</b>) and especially (<b>d</b>) where the diameter is of the level of 100 nm and below.</p>
Full article ">Figure 15
<p>Raman spectrum of HH5 where the D, G, and 2D bands are shown at 1344 cm<sup>−1</sup>, 1600 cm<sup>−1</sup>, and 2714.3 cm<sup>−1</sup>, respectively, and the additional D+G and 2D’ bands at 2907 cm<sup>−1</sup> and 3181 cm<sup>−1</sup>, respectively.</p>
Full article ">
19 pages, 7988 KiB  
Article
Enhancement of Ni-NiO-CeO2 Interaction on Ni–CeO2/Al2O3-MgO Catalyst by Ammonia Vapor Diffusion Impregnation for CO2 Reforming of CH4
by Sabaithip Tungkamani, Saowaluk Intarasiri, Wassachol Sumarasingha, Tanakorn Ratana and Monrudee Phongaksorn
Molecules 2024, 29(12), 2803; https://doi.org/10.3390/molecules29122803 - 12 Jun 2024
Viewed by 579
Abstract
Ni-based catalysts have been widely used for the CO2 reforming of methane (CRM) process, but deactivation is their main problem. This study created an alternative electronic Ni-NiO-CeO2 interaction on the surface of 5 wt% Ni-5 wt% CeO2/Al2O [...] Read more.
Ni-based catalysts have been widely used for the CO2 reforming of methane (CRM) process, but deactivation is their main problem. This study created an alternative electronic Ni-NiO-CeO2 interaction on the surface of 5 wt% Ni-5 wt% CeO2/Al2O3-MgO (5Ni5Ce(xh)/MA) catalysts to enhance catalytic potential simultaneously with coke resistance for the CRM process. The Ni-NiO-CeO2 network was developed on Al2O3-MgO through layered double hydroxide synthesis via our ammonia vapor diffusion impregnation method. The physical properties of the fresh catalysts were analyzed employing FESEM, N2 physisorption, and XRD. The chemical properties on the catalyst surface were analyzed employing H2-TPR, XPS, H2-TPD, CO2-TPD, and O2-TPD. The CRM performances of reduced catalysts were evaluated at 600 °C under ambient pressure. Carbon deposits on spent catalysts were determined quantitatively and qualitatively by TPO, FESEM, and XRD. Compared to 5 wt% Ni-5 wt% CeO2/Al2O3-MgO prepared by the traditional impregnation method, the electronic interaction of the Ni-NiO-CeO2 network with the Al2O3-MgO support was constructed along the time of ammonia diffusion treatment. The electronic interaction in the Ni-NiO-CeO2 nanostructure of the treated catalyst develops surface hydroxyl sites with an efficient pathway of OH* and O* transfer that improves catalytic activities and coke oxidation. Full article
Show Figures

Figure 1

Figure 1
<p>FESEM micrographs of calcined (<b>a</b>) 5Ni5Ce/MA, (<b>b</b>) 5Ni5Ce(6 h)/MA, and (<b>c</b>) 5Ni5Ce(20 h)/MA catalysts.</p>
Full article ">Figure 2
<p>N<sub>2</sub> adsorption–desorption isotherms of support and all calcined catalysts.</p>
Full article ">Figure 3
<p>BJH pore size distributions of support and all calcined catalysts.</p>
Full article ">Figure 4
<p>The X-ray diffraction patterns of (<b>a</b>) MA support and calcined catalysts, and (<b>b</b>) reduced catalysts.</p>
Full article ">Figure 5
<p>H<sub>2</sub>-TPR profiles of all reduced catalysts.</p>
Full article ">Figure 6
<p>XPS core level spectra of reduced catalysts in (<b>a</b>) Ni 2p + Ce 3d energy region and (<b>b</b>) O 1s energy region.</p>
Full article ">Figure 7
<p>H<sub>2</sub>-TPD profiles of all reduced catalysts.</p>
Full article ">Figure 8
<p>CO<sub>2</sub>-TPD profiles of all reduced catalysts.</p>
Full article ">Figure 9
<p>O<sub>2</sub>-TPD profiles of all reduced catalysts.</p>
Full article ">Figure 10
<p>(<b>a</b>) CH<sub>4</sub> conversion, (<b>b</b>) CO<sub>2</sub> conversion, (<b>c</b>) H<sub>2</sub> yield, (<b>d</b>) CO yield, and (<b>e</b>) H<sub>2</sub>/CO ratio for the CO<sub>2</sub> reforming of methane (CRM) of all catalysts. Reaction conditions: 600 °C and 1 atm for 16 h.</p>
Full article ">Figure 11
<p>TPO profile of spent catalysts after CRM reaction.</p>
Full article ">Figure 12
<p>FESEM micrographs of spent (<b>a</b>) 5Ni5Ce/MA, (<b>b</b>) 5Ni5Ce(6 h)/MA, and (<b>c</b>) 5Ni5Ce(20 h)/MA catalysts.</p>
Full article ">Figure 13
<p>The X-ray diffraction patterns of the reduced catalysts and spent catalysts.</p>
Full article ">
17 pages, 5246 KiB  
Article
Nanoimprint Lithography for Next-Generation Carbon Nanotube-Based Devices
by Svitlana Fialkova, Sergey Yarmolenko, Arvind Krishnaswamy, Jagannathan Sankar, Vesselin Shanov, Mark J. Schulz and Salil Desai
Nanomaterials 2024, 14(12), 1011; https://doi.org/10.3390/nano14121011 - 11 Jun 2024
Viewed by 871
Abstract
This research reports the development of 3D carbon nanostructures that can provide unique capabilities for manufacturing carbon nanotube (CNT) electronic components, electrochemical probes, biosensors, and tissue scaffolds. The shaped CNT arrays were grown on patterned catalytic substrate by chemical vapor deposition (CVD) method. [...] Read more.
This research reports the development of 3D carbon nanostructures that can provide unique capabilities for manufacturing carbon nanotube (CNT) electronic components, electrochemical probes, biosensors, and tissue scaffolds. The shaped CNT arrays were grown on patterned catalytic substrate by chemical vapor deposition (CVD) method. The new fabrication process for catalyst patterning based on combination of nanoimprint lithography (NIL), magnetron sputtering, and reactive etching techniques was studied. The optimal process parameters for each technique were evaluated. The catalyst was made by deposition of Fe and Co nanoparticles over an alumina support layer on a Si/SiO2 substrate. The metal particles were deposited using direct current (DC) magnetron sputtering technique, with a particle ranging from 6 nm to 12 nm and density from 70 to 1000 particles/micron. The Alumina layer was deposited by radio frequency (RF) and reactive pulsed DC sputtering, and the effect of sputtering parameters on surface roughness was studied. The pattern was developed by thermal NIL using Si master-molds with PMMA and NRX1025 polymers as thermal resists. Catalyst patterns of lines, dots, and holes ranging from 70 nm to 500 nm were produced and characterized by scanning electron microscopy (SEM) and atomic force microscopy (AFM). Vertically aligned CNTs were successfully grown on patterned catalyst and their quality was evaluated by SEM and micro-Raman. The results confirm that the new fabrication process has the ability to control the size and shape of CNT arrays with superior quality. Full article
(This article belongs to the Section 2D and Carbon Nanomaterials)
Show Figures

Figure 1

Figure 1
<p>Catalyst patterning processes: (<b>A</b>) “negative” and (<b>B</b>) “positive” pattern replica/transfer.</p>
Full article ">Figure 2
<p>SEM images of imprinted substrate (<b>a</b>) and (<b>b</b>) catalyst pattern (Process A).</p>
Full article ">Figure 3
<p>Imprinted substrate (<b>a</b>) and (<b>b</b>) catalyst pattern (Process B).</p>
Full article ">Figure 4
<p>Spin-curve for NRX1025 (<b>a</b>) 2.5% and (<b>b</b>) 7% solution.</p>
Full article ">Figure 5
<p>Spin-curve for PMMA mr-I 35k thermal resist.</p>
Full article ">Figure 6
<p>AFM study of catalyst deposition. DC sputtering at working pressures: (<b>a</b>) 1 mTorr, (<b>b</b>) 2 mTorr, (<b>c</b>) 4 mTorr, and (<b>d</b>) 6 mTorr, respectively.</p>
Full article ">Figure 7
<p>Plasma etching rates for NRX1025: (<b>a</b>) effect of plasma power, (<b>b</b>) effect of oxygen content.</p>
Full article ">Figure 8
<p>Plasma etching rates for PMMA.</p>
Full article ">Figure 9
<p>AFM study of plasma etching effect on thermal resist.</p>
Full article ">Figure 10
<p>SEM study of plasma etching effect: (<b>a</b>) imprinted; (<b>b</b>) etched for 5 min; (<b>c</b>) etched for 8 min.</p>
Full article ">Figure 11
<p>SEM images of short CNTs arrays fabricated by first approach (Process A) with the catalyst patterned by stamp with holes of 290 nm diameter.</p>
Full article ">Figure 12
<p>SEM images of short CNTs arrays fabricated by first approach (Process A) with the catalyst patterned by stamp with lines of 500 nm width.</p>
Full article ">Figure 13
<p>SEM images of short CNTs arrays fabricated by a second approach (Process B) with the catalyst patterned by stamp with dots of 160 nm diameter.</p>
Full article ">Figure 14
<p>SEM images of long CNTs arrays fabricated by the first approach (Process A) with the catalyst patterned by stamp with lines of 70 nm width, 140 nm pattern period. Top view of a CNT arrays at high (<b>a</b>) and low (<b>b</b>) magnifications. Catalyst patterns under CNT arrays (<b>c</b>) SEM image and (<b>d</b>) AFM topography map.</p>
Full article ">Figure 15
<p>SEM images of long CNTs arrays fabricated by the second approach (Process B) with the catalyst patterned by stamp with lines of 455 nm width, 843 nm pattern period. Top view of a CNT arrays at (<b>a</b>) low and <b>(b</b>) high magnifications.</p>
Full article ">Figure 16
<p>HR-SEM images of individual CNTs in array at (<b>a</b>) 200k and (<b>b</b>) 500k magnifications.</p>
Full article ">Figure 17
<p>Raman spectra of reference and patterned CNT arrays.</p>
Full article ">
13 pages, 5815 KiB  
Article
Synthesis, Characterization, and Photocatalytic Properties of Sol-Gel Ce-TiO2 Films
by Lidija Ćurković, Debora Briševac, Davor Ljubas, Vilko Mandić and Ivana Gabelica
Processes 2024, 12(6), 1144; https://doi.org/10.3390/pr12061144 - 1 Jun 2024
Cited by 1 | Viewed by 725
Abstract
In this study, nanostructured cerium-doped TiO2 (Ce-TiO2) films with the addition of different amounts of cerium (0.00, 0.08, 0.40, 0.80, 2.40, and 4.10 wt.%) were deposited on a borosilicate glass substrate by the flow coating sol-gel process. After flow coating, [...] Read more.
In this study, nanostructured cerium-doped TiO2 (Ce-TiO2) films with the addition of different amounts of cerium (0.00, 0.08, 0.40, 0.80, 2.40, and 4.10 wt.%) were deposited on a borosilicate glass substrate by the flow coating sol-gel process. After flow coating, the deposited films were dried at a temperature of 100 °C for 1 h, followed by calcination at a temperature of 450 °C for 2 h. For the characterization of sol-gel TiO2 films, the following analytic techniques were used: X-ray diffraction (XRD), differential thermal analysis (DTA), thermal gravimetry (TG), differential scanning calorimetry (DSC), diffuse reflectance spectroscopy (DRS), and energy dispersive X-ray spectroscopy (EDS). Sol-gel-derived Ce-TiO2 films were used for photocatalytic degradation of ciprofloxacin (CIP). The influence of the amount of Ce in TiO2 films, the duration of the photocatalytic decomposition, and the irradiation type (UV-A and simulated solar light) on the CIP degradation were monitored. Kinetics parameters (reaction kinetics constants and the half-life) of the CIP degradation, as well as photocatalytic degradation efficiency, were determined. The best photocatalytic activity was achieved by the TiO2 film doped with 0.08 wt.% Ce, under both UV-A and solar irradiation. The immobilized catalyst was successfully reused for three cycles under solar light simulator radiation, with changes in photocatalytic efficiency below 3%. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart for the preparation of Ce-TiO<sub>2</sub> films by the sol-gel flow coating method.</p>
Full article ">Figure 2
<p>(<b>A</b>) Side- and (<b>B</b>) cross-section view of the photoreactor.</p>
Full article ">Figure 3
<p>DSC/TG curves for TiO<sub>2</sub> without the addition of Ce (purple line) and TiO<sub>2</sub> with the addition of 0.08 wt.% Ce (blue line).</p>
Full article ">Figure 4
<p>X-ray diffraction patterns and crystallite size TiO<sub>2</sub> samples with different amounts (wt.%) of Ce: 0.00, 0.08, 0.40, 0.80, 2.40, and 4.10.</p>
Full article ">Figure 5
<p>Parameters and volume of unit cells of TiO<sub>2</sub> samples with different amounts (wt.%) of Ce: 0.00, 0.08, 0.40, 0.80, 2.40, and 4.10.</p>
Full article ">Figure 6
<p>(<b>A</b>) DRS Kubelka–Munk spectra for TiO<sub>2</sub> and all Ce-doped samples, and (<b>B</b>) Tauc’s plots with bandgap energy values.</p>
Full article ">Figure 7
<p>Energy-dispersive X-ray (EDS) spectrum of the TiO<sub>2</sub> and all Ce-doped samples.</p>
Full article ">Figure 8
<p>Photolytic and photocatalytic degradation of CIP under (<b>A</b>) UV-A (365 nm) and (<b>B</b>) solar light simulator radiation by sol-gel Ce-TiO<sub>2</sub> films as a function of irradiation time, <span class="html-italic">γ</span><sub>0</sub> (CIP) = 5 mg·L<sup>−1</sup>.</p>
Full article ">Figure 9
<p>Linear transformation of −ln (<span class="html-italic">A</span>/<span class="html-italic">A</span><sub>0</sub>) versus <span class="html-italic">t</span> of photolytic and photocatalytic degradation of CIP under (<b>A</b>) UV-A (365 nm) and (<b>B</b>) solar light simulator radiation by sol-gel Ce-TiO<sub>2</sub> films, <span class="html-italic">γ</span><sub>0</sub> (CIP) = 5 mg·L<sup>−1</sup>.</p>
Full article ">Figure 10
<p>Comparison of the values of (<b>A</b>) pseudo-first-order kinetic constants and (<b>B</b>) degradation efficiency of photolytic and photocatalytic degradation of CIP under UV-A (365 nm) and solar light simulator radiation by sol-gel Ce-TiO<sub>2</sub> films, <span class="html-italic">γ</span><sub>0</sub> (CIP) = 5 mg·L<sup>−1</sup>.</p>
Full article ">Figure 11
<p>Reusability efficiency (<span class="html-italic">η</span>, %) of the sol-gel Ce-TiO<sub>2</sub> films for three consecutive cycles under solar light simulator radiation for TiO<sub>2</sub> film with different amounts of Ce: (<b>A</b>) 0.00 wt.%, (<b>B</b>) 0.08 wt.%, (<b>C</b>) 0.40 wt.%, (<b>D</b>) 0.80 wt.%, (<b>E</b>) 2.40 wt.%, and (<b>F</b>) 4.10 wt.%.</p>
Full article ">
19 pages, 7062 KiB  
Article
Formation of Black Silicon in a Process of Plasma Etching with Passivation in a SF6/O2 Gas Mixture
by Andrey Miakonkikh and Vitaly Kuzmenko
Nanomaterials 2024, 14(11), 945; https://doi.org/10.3390/nano14110945 - 28 May 2024
Viewed by 571
Abstract
This article discusses a method for forming black silicon using plasma etching at a sample temperature range from −20 °C to +20 °C in a mixture of oxygen and sulfur hexafluoride. The surface morphology of the resulting structures, the autocorrelation function of surface [...] Read more.
This article discusses a method for forming black silicon using plasma etching at a sample temperature range from −20 °C to +20 °C in a mixture of oxygen and sulfur hexafluoride. The surface morphology of the resulting structures, the autocorrelation function of surface features, and reflectivity were studied depending on the process parameters—the composition of the plasma mixture, temperature and other discharge parameters (radical concentrations). The relationship between these parameters and the concentrations of oxygen and fluorine radicals in plasma is shown. A novel approach has been studied to reduce the reflectance using conformal bilayer dielectric coatings deposited by atomic layer deposition. The reflectivity of the resulting black silicon was studied in a wide spectral range from 400 to 900 nm. As a result of the research, technologies for creating black silicon on silicon wafers with a diameter of 200 mm have been proposed, and the structure formation process takes no more than 5 min. The resulting structures are an example of the self-formation of nanostructures due to anisotropic etching in a gas discharge plasma. This material has high mechanical, chemical and thermal stability and can be used as an antireflective coating, in structures requiring a developed surface—photovoltaics, supercapacitors, catalysts, and antibacterial surfaces. Full article
(This article belongs to the Special Issue Synthesis of Nanostructures in Gas-Discharge Plasma)
Show Figures

Figure 1

Figure 1
<p>Example of silicon nanograss formation in continuous cryogenic deep silicon etching process in mix of SF<sub>6</sub> and O<sub>2</sub> with over-passivation.</p>
Full article ">Figure 2
<p>The example of autocorrelation function: (<b>a</b>) SEM picture of top view of black silicon; (<b>b</b>) radial autocorrelation function for that image.</p>
Full article ">Figure 3
<p>The needles height or holes depth over SF<sub>6</sub> fraction in plasma mixture, duration of etch is 300 s, temperature is −20 °C.</p>
Full article ">Figure 4
<p>Schematic representation of the formation of black silicon at high oxygen content in the feed mixture (<b>left</b>) and low oxygen content in the feed mixture (<b>right</b>). In this figure, positive ions are shown as circles with a plus sign. F* and O* stand for fluorine and oxygen radicals, respectively.</p>
Full article ">Figure 5
<p>SEM pictures of black silicon for different duration of etching ((<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>)—side view, (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>)—top view): (<b>a</b>,<b>b</b>) 60 s; (<b>c</b>,<b>d</b>) 120 s; (<b>e</b>,<b>f</b>) 300 s; (<b>g</b>,<b>h</b>) 600 s. For all processes the same set of parameters was chosen—temperature is −20 °C, fraction of SF<sub>6</sub> is 0.46.</p>
Full article ">Figure 6
<p>Black silicon characteristics over etching duration: (<b>a</b>) needles height or holes depth; (<b>b</b>) autocorrelation length. For all processes the same set of parameters was chosen—temperature is −20 °C, fraction of SF<sub>6</sub> is 0.46.</p>
Full article ">Figure 7
<p>SEM picture of black silicon for different DC bias voltage during etching process: (<b>a</b>) 160 V; (<b>b</b>) 58 V. For both processes the same set of parameters was chosen—temperature is −20 °C, fraction of SF<sub>6</sub> is 0.46.</p>
Full article ">Figure 8
<p>The spectral reflectivity coefficient of black silicon for different parameters of etching process: (<b>a</b>) etching duration (temperature is −20 °C, fraction of SF<sub>6</sub> (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ν</mi> </mrow> <mrow> <msub> <mrow> <mi>S</mi> <mi>F</mi> </mrow> <mrow> <mn>6</mn> </mrow> </msub> </mrow> </msub> </mrow> </semantics></math>) is 0.46); (<b>b</b>) SF<sub>6</sub> fraction in plasma mixture (temperature is −20 °C, etching duration is 300 s); (<b>c</b>) wafer temperature (fraction of SF<sub>6</sub> is 0.46, etching duration is 300 s).</p>
Full article ">Figure 9
<p>Modeled spectral reflectivity coefficient of plane two-layer antireflective coating: (<b>a</b>) Si(bulk)/Al<sub>2</sub>O<sub>3</sub>(30 nm)/HfO<sub>2</sub>(different thickness); (<b>b</b>) Si(bulk)/Al<sub>2</sub>O<sub>3</sub>(80 nm)/HfO<sub>2</sub>(different thickness).</p>
Full article ">Figure 10
<p>The measured spectral reflectivity coefficient of black silicon with different antireflective coatings. The black silicon etching parameters are as follows—temperature is −20 °C, fraction of SF<sub>6</sub> is 0.46, etching duration is 300 s.</p>
Full article ">Figure 11
<p>Plasma parameters over SF<sub>6</sub> fraction in plasma mixture: (<b>a</b>) electron temperature; (<b>b</b>) electrons and positive ions concentrations; (<b>c</b>) fluorine and oxygen radical concentrations.</p>
Full article ">
22 pages, 2019 KiB  
Review
Ball Milling Innovations Advance Mg-Based Hydrogen Storage Materials Towards Practical Applications
by Yaohui Xu, Yuting Li, Quanhui Hou, Yechen Hao and Zhao Ding
Materials 2024, 17(11), 2510; https://doi.org/10.3390/ma17112510 - 23 May 2024
Cited by 1 | Viewed by 624
Abstract
Mg-based materials have been widely studied as potential hydrogen storage media due to their high theoretical hydrogen capacity, low cost, and abundant reserves. However, the sluggish hydrogen absorption/desorption kinetics and high thermodynamic stability of Mg-based hydrides have hindered their practical application. Ball milling [...] Read more.
Mg-based materials have been widely studied as potential hydrogen storage media due to their high theoretical hydrogen capacity, low cost, and abundant reserves. However, the sluggish hydrogen absorption/desorption kinetics and high thermodynamic stability of Mg-based hydrides have hindered their practical application. Ball milling has emerged as a versatile and effective technique to synthesize and modify nanostructured Mg-based hydrides with enhanced hydrogen storage properties. This review provides a comprehensive summary of the state-of-the-art progress in the ball milling of Mg-based hydrogen storage materials. The synthesis mechanisms, microstructural evolution, and hydrogen storage properties of nanocrystalline and amorphous Mg-based hydrides prepared via ball milling are systematically reviewed. The effects of various catalytic additives, including transition metals, metal oxides, carbon materials, and metal halides, on the kinetics and thermodynamics of Mg-based hydrides are discussed in detail. Furthermore, the strategies for synthesizing nanocomposite Mg-based hydrides via ball milling with other hydrides, MOFs, and carbon scaffolds are highlighted, with an emphasis on the importance of nanoconfinement and interfacial effects. Finally, the challenges and future perspectives of ball-milled Mg-based hydrides for practical on-board hydrogen storage applications are outlined. This review aims to provide valuable insights and guidance for the development of advanced Mg-based hydrogen storage materials with superior performance. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Flowchart for the preparation method of the as-milled Sm<sub>5</sub>Mg4<sub>1</sub> alloy [<a href="#B45-materials-17-02510" class="html-bibr">45</a>]. (<b>b</b>) Illustration of the deformation of powder agglomerate during the impact process [<a href="#B46-materials-17-02510" class="html-bibr">46</a>]. (<b>c</b>) Illustration of the different microstructural states of the Mg<sub>2</sub>Ni alloys [<a href="#B50-materials-17-02510" class="html-bibr">50</a>].</p>
Full article ">Figure 2
<p>SEM images and schematic models of samples: (<b>a</b>) MgH<sub>2</sub>-mNi (2 h); (<b>b</b>) MgH<sub>2</sub>-nNi (2 h) [<a href="#B80-materials-17-02510" class="html-bibr">80</a>]; (<b>c</b>) total charge densities in (110) crystal planes of Mg<sub>16</sub>H<sub>32</sub>,Mg<sub>14</sub>Ni<sub>2</sub>H<sub>32</sub> [<a href="#B81-materials-17-02510" class="html-bibr">81</a>]; (<b>d</b>) schematic diagram of the de/re-hydrogenation processes of the MgH<sub>2</sub> + NiO@NiMoO<sub>4</sub> composite [<a href="#B82-materials-17-02510" class="html-bibr">82</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) Schematic of the automated ball milling with an aerosol spraying device. (<b>b</b>) Schematic of the dehydrogenation via the MgH<sub>2</sub>-Mg pathway. (<b>c</b>) The reaction takes place at the Mg/LiBH<sub>4</sub> interface, leading to the nucleation and growth of MgB<sub>2</sub> + LiH products (shown inside the dashed box) [<a href="#B93-materials-17-02510" class="html-bibr">93</a>]. (<b>d</b>) Schematic illustration of the synergistic effects of the TiH<sub>1.5</sub>−Mg<sub>2</sub>Ni nanocatalyst [<a href="#B96-materials-17-02510" class="html-bibr">96</a>].</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic diagram for the preparation of 2D MOF@Pd hybrid nanosheets [<a href="#B99-materials-17-02510" class="html-bibr">99</a>]. (<b>b</b>) Schematic illustration of the catalytic mechanism of MgH<sub>2</sub>-5 wt.% Ni MOF [<a href="#B100-materials-17-02510" class="html-bibr">100</a>]. (<b>c</b>) Schematic illustration of catalytic mechanism of the MgH<sub>2</sub>-TM MOF (TM = Fe, Ni) composite [<a href="#B101-materials-17-02510" class="html-bibr">101</a>].</p>
Full article ">
24 pages, 1120 KiB  
Review
Recent Advances in the Preparation Methods of Magnesium-Based Hydrogen Storage Materials
by Yaohui Xu, Yang Zhou, Yuting Li, Yechen Hao, Pingkeng Wu and Zhao Ding
Molecules 2024, 29(11), 2451; https://doi.org/10.3390/molecules29112451 - 23 May 2024
Cited by 2 | Viewed by 967
Abstract
Magnesium-based hydrogen storage materials have garnered significant attention due to their high hydrogen storage capacity, abundance, and low cost. However, the slow kinetics and high desorption temperature of magnesium hydride hinder its practical application. Various preparation methods have been developed to improve the [...] Read more.
Magnesium-based hydrogen storage materials have garnered significant attention due to their high hydrogen storage capacity, abundance, and low cost. However, the slow kinetics and high desorption temperature of magnesium hydride hinder its practical application. Various preparation methods have been developed to improve the hydrogen storage properties of magnesium-based materials. This review comprehensively summarizes the recent advances in the preparation methods of magnesium-based hydrogen storage materials, including mechanical ball milling, methanol-wrapped chemical vapor deposition, plasma-assisted ball milling, organic ligand-assisted synthesis, and other emerging methods. The principles, processes, key parameters, and modification strategies of each method are discussed in detail, along with representative research cases. Furthermore, the advantages and disadvantages of different preparation methods are compared and evaluated, and their influence on hydrogen storage properties is analyzed. The practical application potential of these methods is also assessed, considering factors such as hydrogen storage performance, scalability, and cost-effectiveness. Finally, the existing challenges and future research directions in this field are outlined, emphasizing the need for further development of high-performance and cost-effective magnesium-based hydrogen storage materials for clean energy applications. This review provides valuable insights and references for researchers working on the development of advanced magnesium-based hydrogen storage technologies. Full article
(This article belongs to the Special Issue Recent Advances of Hydrogen Storage Hydride Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>–<b>c</b>) Illustration of the deformation of powder agglomerate during the impact process [<a href="#B28-molecules-29-02451" class="html-bibr">28</a>]; strong external magnets induce (<b>d</b>) high-energy shearing (HES), (<b>e</b>) intense impact (IMP2) milling modes [<a href="#B30-molecules-29-02451" class="html-bibr">30</a>].</p>
Full article ">Figure 2
<p>Schematic illustration of the MWCVD process [<a href="#B49-molecules-29-02451" class="html-bibr">49</a>].</p>
Full article ">Figure 3
<p>Schematic illustration of the preparation process of Mg<sub>85</sub>In<sub>5</sub>Al<sub>5</sub>Ti<sub>5</sub> alloy by P-milling [<a href="#B65-molecules-29-02451" class="html-bibr">65</a>].</p>
Full article ">Figure 4
<p>Synthesis procedure of composite through organic ligand-assisted strategy [<a href="#B72-molecules-29-02451" class="html-bibr">72</a>].</p>
Full article ">
34 pages, 4088 KiB  
Review
Magnetic Iron Oxide Nanomaterials for Lipase Immobilization: Promising Industrial Catalysts for Biodiesel Production
by Farid Hajareh Haghighi, Roya Binaymotlagh, Cleofe Palocci and Laura Chronopoulou
Catalysts 2024, 14(6), 336; https://doi.org/10.3390/catal14060336 - 22 May 2024
Viewed by 791
Abstract
Biodiesel is a mixture of fatty acid alkyl esters (FAAEs) mainly produced via transesterification reactions among triglycerides and short-chain alcohols catalyzed by chemical catalysts (e.g., KOH, NaOH). Lipase-assisted enzymatic transesterification has been proposed to overcome the drawbacks of chemical synthesis, such as high [...] Read more.
Biodiesel is a mixture of fatty acid alkyl esters (FAAEs) mainly produced via transesterification reactions among triglycerides and short-chain alcohols catalyzed by chemical catalysts (e.g., KOH, NaOH). Lipase-assisted enzymatic transesterification has been proposed to overcome the drawbacks of chemical synthesis, such as high energy consumption, expensive separation of the catalyst from the reaction mixture and production of large amounts of wastewater during product separation and purification. However, one of the main drawbacks of this process is the enzyme cost. In recent years, nano-immobilized lipases have received extensive attention in the design of robust industrial biocatalysts for biodiesel production. To improve lipase catalytic efficiency, magnetic nanoparticles (MNPs) have attracted growing interest as versatile lipase carriers, owing to their unique properties, such as high surface-to-volume ratio and high enzyme loading capacity, low cost and inertness against chemical and microbial degradation, biocompatibility and eco-friendliness, standard synthetic methods for large-scale production and, most importantly, magnetic properties, which provide the possibility for the immobilized lipase to be easily separated at the end of the process by applying an external magnetic field. For the preparation of such effective magnetic nano-supports, various surface functionalization approaches have been developed to immobilize a broad range of industrially important lipases. Immobilization generally improves lipase chemical-thermal stability in a wide pH and temperature range and may also modify its catalytic performance. Additionally, different lipases can be co-immobilized onto the same nano-carrier, which is a highly effective strategy to enhance biodiesel yield, specifically for those feedstocks containing heterogeneous free fatty acids (FFAs). This review will present an update on the use of magnetic iron oxide nanostructures (MNPs) for lipase immobilization to catalyze transesterification reactions for biodiesel production. The following aspects will be covered: (1) common organic modifiers for magnetic nanoparticle support and (2) recent studies on modified MNPs-lipase catalysts for biodiesel production. Aspects concerning immobilization procedures and surface functionalization of the nano-supports will be highlighted. Additionally, the main features that characterize these nano-biocatalysts, such as enzymatic activity, reusability, resistance to heat and pH, will be discussed. Perspectives and key considerations for optimizing biodiesel production in terms of sustainability are also provided for future studies. Full article
Show Figures

Figure 1

Figure 1
<p>Enzyme-based transesterification reaction. Reprinted from Ref. [<a href="#B18-catalysts-14-00336" class="html-bibr">18</a>], copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 2
<p>Main types of enzyme immobilization strategies. Adapted from Ref. [<a href="#B29-catalysts-14-00336" class="html-bibr">29</a>], copyright 2022, with permission from Elsevier.</p>
Full article ">Figure 3
<p>Transesterification of animal fat to biodiesel. TGL: triacylglycerol lipase; nsTGL: non specific triacylglycerol lipase; MGL: monoacylglycerol lipase. Reprinted from Ref. [<a href="#B85-catalysts-14-00336" class="html-bibr">85</a>], MDPI, 2020.</p>
Full article ">Figure 4
<p>Biodiesel production using: (<b>a</b>) alkali-catalyzed and (<b>b</b>) enzyme-catalyzed transesterification processes. Adapted from Ref. [<a href="#B18-catalysts-14-00336" class="html-bibr">18</a>], copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 5
<p>Bioreactors developed for enzymatic biodiesel production: stirred tank reactor (<b>a</b>), packed-bed reactor (<b>b</b>) and fluidized-bed reactor (<b>c</b>). Reprinted from Ref. [<a href="#B18-catalysts-14-00336" class="html-bibr">18</a>], copyright 2020, with permission from Elsevier.</p>
Full article ">Figure 6
<p>Schematic routes for the preparation of Fe<sub>3</sub>O<sub>4</sub>NPs@TEOS-TSD nanocomposite and immobilization of CALB onto the magnetic support. Adapted from Ref. [<a href="#B156-catalysts-14-00336" class="html-bibr">156</a>], copyright (2021), with permission from Elsevier.</p>
Full article ">Figure 7
<p>Preparation of magnetic Fe<sub>3</sub>O<sub>4</sub>NPs-APTES-GA-RML. Redrawn from Ref. [<a href="#B164-catalysts-14-00336" class="html-bibr">164</a>], copyright 2022, with permission from Springer.</p>
Full article ">Figure 8
<p>Schematic representation of the cyclic process for the biocatalytic synthesis of biodiesel in ILs. Adapted from Ref. [<a href="#B164-catalysts-14-00336" class="html-bibr">164</a>], copyright (2022), with permission from Springer.</p>
Full article ">Figure 9
<p>Schematic routes for the preparation of biocatalysts to produce biodiesel. Redrawn from Ref. [<a href="#B173-catalysts-14-00336" class="html-bibr">173</a>], copyright 2020, with permission from Elsevier.</p>
Full article ">Figure 10
<p>Schematic routes for the preparation of biocatalysts to produce biodiesel. Adapted from Ref. [<a href="#B179-catalysts-14-00336" class="html-bibr">179</a>], copyright (2020), with permission from Elsevier.</p>
Full article ">Figure 11
<p>Schematic illustration of the preparation process of magnetic Fe<sub>3</sub>O<sub>4</sub>NPs-Ti-GO-CALB biocatalyst. Redrawn from Ref. [<a href="#B180-catalysts-14-00336" class="html-bibr">180</a>], copyright 2023, with permission from Elsevier.</p>
Full article ">Figure 12
<p>Schematic illustration of the preparation process for magnetic activated lipase-inorganic hybrid nanoflowers (MhNF). Adapted from Ref. [<a href="#B184-catalysts-14-00336" class="html-bibr">184</a>], copyright 2021, with permission from Elsevier.</p>
Full article ">Figure 13
<p>Schematic diagram of the preparation of co-immobilized Tween 80-AOL@RML. (Key: AOL: free <span class="html-italic">Aspergillus oryzae</span> lipase; RML: free <span class="html-italic">Rhizomucor miehei</span> lipase; Tween-AOL: <span class="html-italic">Aspergillus oryzae</span> lipase bio-imprinted using Tween 80; Tween-RML: <span class="html-italic">Rhizomucor miehei</span> lipase bio-imprinted using Tween 80; co-im Tween-AOL@RML: co-immobilized Tween-AOL and Tween-RML). Adapted from Ref. [<a href="#B197-catalysts-14-00336" class="html-bibr">197</a>], copyright 2022, with permission from Elsevier.</p>
Full article ">Figure 14
<p>Schematic representation of lipase immobilization on dopamine-coated MNPs. Redrawn from Ref. [<a href="#B200-catalysts-14-00336" class="html-bibr">200</a>] under the terms and conditions of the Creative Commons Attribution (CC BY) license, MDPI.</p>
Full article ">Figure 15
<p>Preparation of Fe<sub>3</sub>O<sub>4</sub>-poly(GMA-<span class="html-italic">co</span>-MAA) composite and subsequent immobilization of lipase onto the magnetic support. Abbreviations: NHS = N-hydroxysulfosuccinimide; EDC = 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide;. GMA = glycidyl methahacrylate; DVB = divinylbenezene; PAAS = sodium polyacrylate; MAA, methacrylic acid. Adapted from Ref. [<a href="#B203-catalysts-14-00336" class="html-bibr">203</a>], copyright 2020, with permission from Elsevier.</p>
Full article ">
12 pages, 5162 KiB  
Article
Synthesis and Mechanism Study of Carbon Nanowires, Carbon Nanotubes, and Carbon Pompons on Single-Crystal Diamonds
by Shuai Wu, Qiang Wang, Kesheng Guo, Lei Liu, Jie Bai, Zhenhuai Yang, Xin Li and Hong Liu
Crystals 2024, 14(6), 481; https://doi.org/10.3390/cryst14060481 - 21 May 2024
Cited by 1 | Viewed by 627
Abstract
Carbon nanomaterials are in high demand owing to their exceptional physical and chemical properties. This study employed a mixture of CH4, H2, and N2 to create carbon nanostructures on a single-crystal diamond using microwave plasma chemical vapor deposition [...] Read more.
Carbon nanomaterials are in high demand owing to their exceptional physical and chemical properties. This study employed a mixture of CH4, H2, and N2 to create carbon nanostructures on a single-crystal diamond using microwave plasma chemical vapor deposition (MPCVD) under high-power conditions. By controlling the substrate surface and nitrogen flow rate, carbon nanowires, carbon nanotubes, and carbon pompons could be selectively deposited. The results obtained from OES, SEM, TEM, and Raman spectroscopy revealed that the nitrogen flow rate and substrate surface conditions were crucial for the growth of carbon nanostructures. The changes in the plasma shape enhanced the etching effect, promoting the growth of carbon pompons. The CN and C2 groups play vital catalytic roles in the formation of carbon nanotubes and nanowires, guiding the precipitation and composite growth of carbon atoms at the interface between the Mo metal catalysts and diamond. This study demonstrated that heterostructures of diamond–carbon nanomaterials could be produced under high-power conditions, offering a new approach to integrating diamond and carbon nanomaterials. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of single-crystal diamond substrate. (<b>a</b>) Diamond substrates for experiments 1 and 2, with rectangular pits marked by dashed lines in the middle. (<b>b</b>) Diamond substrates for experiments 3 and 4, with molybdenum flakes in black.</p>
Full article ">Figure 2
<p>(<b>a</b>) SEM images of CPs deposited in diamond rectangular pits in experiment 1, (<b>b</b>) SEM images of CNPs deposited on diamond surfaces in experiment 2, (<b>c</b>) SEM images of CNWs deposited in experiment 3, and (<b>d</b>) SEM images of CNTs deposited in experiment 4.</p>
Full article ">Figure 3
<p>TEM images of (<b>a</b>) CPs deposited in experiment 1 and (<b>b</b>) CNWs deposited in experiment 3.</p>
Full article ">Figure 4
<p>Raman spectra of (<b>a</b>) CPs deposited in experiment 1, (<b>b</b>) CNWs deposited in experiment 3, and (<b>c</b>) CNTs deposited in experiment 4.</p>
Full article ">Figure 5
<p>OES spectra of (<b>a</b>) plasma in experiment 1 and (<b>b</b>) plasmons in experiments 3 and 4.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic diagram of CP formation. (<b>b</b>) Schematic diagram of CNW and CNT formation.</p>
Full article ">
Back to TopTop