Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (4,555)

Search Parameters:
Keywords = AlN

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 2990 KiB  
Article
Ni-Mg-Al Hydrotalcite-Derived Catalysts for Ammonia Decomposition—From Precursor to Effective Catalyst
by Andrzej Kowalczyk, Martyna Zaryczny, Zofia Piwowarska and Lucjan Chmielarz
Molecules 2025, 30(5), 1052; https://doi.org/10.3390/molecules30051052 - 25 Feb 2025
Abstract
A series of Ni-Mg-Al hydrotalcite-derived mixed metal oxides with different Ni/Mg ratios were prepared by the coprecipitation method followed by calcination at 600 °C. The hydrotalcite-like materials, as well as their calcined forms, were characterized with respect to structure (XRD, UV-Vis DRS), chemical [...] Read more.
A series of Ni-Mg-Al hydrotalcite-derived mixed metal oxides with different Ni/Mg ratios were prepared by the coprecipitation method followed by calcination at 600 °C. The hydrotalcite-like materials, as well as their calcined forms, were characterized with respect to structure (XRD, UV-Vis DRS), chemical composition (ICP-OES), textural parameters (low-temperature N2 sorption), dispersion of nickel species (H2-chemisorption) and nickel species reducibility (H2-TPR). Moreover, the process of hydrotalcite-like materials’ thermal transformation to mixed metal oxide systems in air and argon flow was studied by the TG-DTA method. The activity of the studied catalysts in the reaction of ammonia decomposition increased with an increase in nickel content in the samples. It was shown that nickel species incorporated into the Mg-Al oxide matrix segregated under conditions of reduction in a flow of H2/Ar mixture with the formation of metallic nickel crystallites of the average size of about 10 nm. The size of nickel crystallites was practically no change in the subsequent reduction cycles and resulted in increased catalytic activity in comparison to larger crystallites of metallic nickel (20.2–23.6 nm) deposited on Al2O3 and MgO. Full article
Show Figures

Figure 1

Figure 1
<p>Diffractograms recorded for the Ni-Mg-Al hydrotalcite-like materials.</p>
Full article ">Figure 2
<p>DTG profiles obtained for the thermal decomposition of hydrotalcite-like materials in flow or air (solid line) and nitrogen (dashed line).</p>
Full article ">Figure 3
<p>Diffractograms of the samples calcined at 600 °C (<b>A</b>) and 800 °C (<b>B</b>) as well as after catalytic test (<b>C</b>).</p>
Full article ">Figure 3 Cont.
<p>Diffractograms of the samples calcined at 600 °C (<b>A</b>) and 800 °C (<b>B</b>) as well as after catalytic test (<b>C</b>).</p>
Full article ">Figure 4
<p>UV-Vis DR spectra recorded for the samples calcined at 600 and 800 °C.</p>
Full article ">Figure 5
<p>H<sub>2</sub>-TPR profiles of the samples calcined at 600 °C (<b>A</b>) and reduction profiles of the MZ02 sample recorded in the subsequent H2-TPR runs (<b>B</b>).</p>
Full article ">Figure 5 Cont.
<p>H<sub>2</sub>-TPR profiles of the samples calcined at 600 °C (<b>A</b>) and reduction profiles of the MZ02 sample recorded in the subsequent H2-TPR runs (<b>B</b>).</p>
Full article ">Figure 6
<p>Diffractograms of the freshly calcined MZ02 sample at 600 °C as well as after catalytic tests and seven H<sub>2</sub>-TPR cycles.</p>
Full article ">Figure 7
<p>Results of the catalytic tests of the ammonia decomposition reaction.</p>
Full article ">Figure 8
<p>Studies of the catalyst’s stability in the subsequent catalytic cycles of ammonia decomposition reaction in the presence of MZ01 (<b>A</b>), MZ02 (<b>B</b>), and MZ03 (<b>C</b>) catalysts.</p>
Full article ">Figure 9
<p>Long-term isothermal stability test for the MZ03 catalyst at 450 °C.</p>
Full article ">
15 pages, 6947 KiB  
Article
Effects of Intermetallic NiAl Particle Content on Friction and Wear of Spark Plasma-Sintered Alumina Matrix Composites
by Nay Win Khun, Mingyue Huang, Zhong Alan Li, He Zhang, Khiam Aik Khor, Jinglei Yang and Fei Duan
Lubricants 2025, 13(3), 101; https://doi.org/10.3390/lubricants13030101 - 25 Feb 2025
Abstract
The spark plasma sintering (SPS) technology was applied to develop alumina matrix composites (Al2O3MCs) with different nickel-aluminium (NiAl) particle contents of 5–20 wt.% to understand a correlation between their NiAl particle contents and their microstructures, fracture, hardness, friction, and [...] Read more.
The spark plasma sintering (SPS) technology was applied to develop alumina matrix composites (Al2O3MCs) with different nickel-aluminium (NiAl) particle contents of 5–20 wt.% to understand a correlation between their NiAl particle contents and their microstructures, fracture, hardness, friction, and wear. The incorporation of NiAl particles suppressed micrograins and micropores in the microstructures of the Al2O3MCs, which resulted in their improved fracture resistance. Increasing the NiAl particle content from 0 to 20 wt.% gave rise to a 23.9% decrease in the hardness of the Al2O3MCs. The Al2O3MCs had 18.2% and 13.3% decreases in their friction coefficients and 68.3% and 81.3% decreases in their specific wear rates under the normal loads of 2 and 6 N, respectively, with an increased NiAl particle content from 0 to 20 wt.% thanks to their decreased fatigue wear. The SPS Al2O3MCs with NiAl particles had promising tribological performance for rotating gas turbine components. Full article
(This article belongs to the Special Issue Friction and Wear of Ceramics)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XPS (<b>a</b>) Ni 2p and (<b>b</b>) Al 2p peaks of Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles.</p>
Full article ">Figure 2
<p>Microstructures of (<b>a</b>) Al<sub>2</sub>O<sub>3</sub> and (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles.</p>
Full article ">Figure 3
<p>Indents on surfaces of (<b>a</b>) Al<sub>2</sub>O<sub>3</sub> and (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles made under a 19.6 N load.</p>
Full article ">Figure 4
<p>Hardness values of Al<sub>2</sub>O<sub>3</sub> and Al<sub>2</sub>O<sub>3</sub>MCs with different NiAl particle contents.</p>
Full article ">Figure 5
<p>Friction coefficients of Al<sub>2</sub>O<sub>3</sub>, Al<sub>2</sub>O<sub>3</sub>MCs with different NiAl particle contents, and NiAl tested under (<b>a</b>) 2 and (<b>b</b>) 6 N loads, as a function of the number of laps.</p>
Full article ">Figure 6
<p>Friction coefficients of Al<sub>2</sub>O<sub>3</sub> and Al<sub>2</sub>O<sub>3</sub>MCs with different NiAl particle contents tested under 2 and 6 N loads.</p>
Full article ">Figure 7
<p>Specific wear rates of Al<sub>2</sub>O<sub>3</sub> and Al<sub>2</sub>O<sub>3</sub>MCs with different NiAl particle contents tested under 2 and 6 N loads.</p>
Full article ">Figure 8
<p>Wear topographies of (<b>a</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles, and (<b>c</b>) NiAl tested under a 6 N load.</p>
Full article ">Figure 9
<p>Wear morphologies of NiAl, tested under (<b>a</b>,<b>c</b>) 2 and (<b>b</b>,<b>d</b>) 6 N loads, observed at different magnifications.</p>
Full article ">Figure 10
<p>Wear morphologies of (<b>a</b>,<b>c</b>,<b>e</b>) Al<sub>2</sub>O<sub>3</sub> and (<b>b</b>,<b>d</b>,<b>f</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles tested under (<b>a</b>,<b>b</b>) 2 and (<b>c</b>–<b>f</b>) 6 N loads.</p>
Full article ">Figure 11
<p>EDX spectra of (<b>a</b>,<b>b</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>,<b>d</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles, and (<b>e</b>,<b>f</b>) NiAl tested under a 6 N load, measured at locations (<b>a</b>) ‘A’ in untested area and (<b>b</b>) ‘B’ in wear track of <a href="#lubricants-13-00101-f010" class="html-fig">Figure 10</a>c, (<b>c</b>) ‘C’ in untested area and (<b>d</b>) ‘D’ in wear track of <a href="#lubricants-13-00101-f010" class="html-fig">Figure 10</a>d, and (<b>e</b>) ‘E’ in untested area and (<b>f</b>) ‘F’ in wear track of <a href="#lubricants-13-00101-f009" class="html-fig">Figure 9</a>b.</p>
Full article ">Figure 11 Cont.
<p>EDX spectra of (<b>a</b>,<b>b</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>,<b>d</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles, and (<b>e</b>,<b>f</b>) NiAl tested under a 6 N load, measured at locations (<b>a</b>) ‘A’ in untested area and (<b>b</b>) ‘B’ in wear track of <a href="#lubricants-13-00101-f010" class="html-fig">Figure 10</a>c, (<b>c</b>) ‘C’ in untested area and (<b>d</b>) ‘D’ in wear track of <a href="#lubricants-13-00101-f010" class="html-fig">Figure 10</a>d, and (<b>e</b>) ‘E’ in untested area and (<b>f</b>) ‘F’ in wear track of <a href="#lubricants-13-00101-f009" class="html-fig">Figure 9</a>b.</p>
Full article ">Figure 12
<p>Wear morphologies of Al<sub>2</sub>O<sub>3</sub> balls slid on (<b>a</b>) Al<sub>2</sub>O<sub>3</sub>, (<b>b</b>) Al<sub>2</sub>O<sub>3</sub>MC with 20 wt.% NiAl particles, and (<b>c</b>) NiAl tested under a6 N load. The scale bars of 100 µm, 100 µm, and 200 µm are used in (<b>a</b>–<b>c</b>), respectively.</p>
Full article ">
15 pages, 4549 KiB  
Article
Performance Analysis of Scandium-Doped Aluminum Nitride-Based PMUTs Under High-Temperature Conditions
by Haochen Lyu and Ahmad Safari
Appl. Sci. 2025, 15(5), 2428; https://doi.org/10.3390/app15052428 - 24 Feb 2025
Abstract
PMUTs have been widely studied in recent years, particularly those based on the SOI (silicon-on-insulator) process, which have been partially commercialized and are extensively used in advanced applications such as ultrasonic ranging and spatial positioning. However, there has been little research on their [...] Read more.
PMUTs have been widely studied in recent years, particularly those based on the SOI (silicon-on-insulator) process, which have been partially commercialized and are extensively used in advanced applications such as ultrasonic ranging and spatial positioning. However, there has been little research on their high-temperature reliability, a critical area for their use in extreme environmental conditions. In this study, we investigate the high-temperature characteristics of air-coupled PMUTs based on SOI under various structural conditions, employing both finite element analysis (FEA) and experimental validation. We assess the performance of PMUTs at elevated temperatures by examining key parameters such as resonant frequency, the electromechanical coupling coefficient, mechanical amplitude, and warpage, all analyzed as functions of temperature. The experimental results show that temperature-induced drift becomes more significant as the back cavity size increases and the top silicon layer thickness decreases. These findings are consistent with the trends observed in the finite element analysis. Specifically, a PMUT with a back cavity diameter of 1000 μm and a top silicon thickness of 4 μm exhibits a temperature drift rate of up to 47.3% when the operating temperature rises from room temperature to 200 °C. Furthermore, at elevated temperatures, the maximum electromechanical coupling coefficient improves by 68.6%, and the mechanical amplitude increases by 66.1%. Heating experiments using a 3D profiler reveal that warpage increases from 0.3 μm to 2.15 μm as the temperature reaches 150 °C. These findings offer important theoretical insights into the temperature-induced drift behavior of PMUTs under high-temperature conditions. This study provides a comprehensive understanding of the performance variations of PMUTs, including changes in electromechanical coupling, mechanical amplitude, and structural warpage, which are critical for their reliable operation in extreme environments. The results presented here can serve as a foundation for the design and optimization of PMUTs in applications that require high-temperature stability, ensuring their enhanced reliability and performance in such demanding conditions. Full article
(This article belongs to the Special Issue Applications of Thin Films and Their Physical Properties)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic of the device structure. (<b>b</b>) Front and back images of two PMUTs with different back cavity sizes.</p>
Full article ">Figure 2
<p>Process flow of the device. (<b>a</b>) BOX wafer. (<b>b</b>) Thin-film deposition. (<b>c</b>) Thin-film etching and metalization. (<b>d</b>) Backside etching.</p>
Full article ">Figure 3
<p>XRD result of AlScN thin film.</p>
Full article ">Figure 4
<p>Frequency variation with temperature for devices with different back cavity sizes (FEM).</p>
Full article ">Figure 5
<p>Frequency variation with temperature for devices with different Si thicknesses (FEM).</p>
Full article ">Figure 6
<p>FEA and experimental result comparison of PMUT with 1000 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m back cavity.</p>
Full article ">Figure 7
<p>Frequency characteristics of four devices with a back cavity size of 1000 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m as temperature changes.</p>
Full article ">Figure 8
<p>Impedance phase variation of four devices with temperature.</p>
Full article ">Figure 9
<p>Variation in electromechanical coupling coefficient of the devices with temperature.</p>
Full article ">Figure 10
<p>Frequency and impedance phase variation with temperature for devices with a 600 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m back cavity (<b>a</b>) Frequency (<b>b</b>) Impedance phase.</p>
Full article ">Figure 11
<p>Frequency variation with temperature for devices with a 600 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m back cavity with different top Si thickness.</p>
Full article ">Figure 12
<p>Amplitude and mechanical resonant frequency variation with temperature for devices with a 1000 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m back cavity (LDV test results).</p>
Full article ">Figure 13
<p>Resonant frequency variation during heating and cooling cycles for devices with different back cavity sizes: (<b>a</b>) 600 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m and (<b>b</b>) 1000 <math display="inline"><semantics> <mi mathvariant="sans-serif">μ</mi> </semantics></math>m.</p>
Full article ">
21 pages, 11884 KiB  
Article
Process Parameters and Heat-Treatment Optimization for Improving Microstructural and Mechanical Properties of AA6082-T651 Deposit on EN14B Plate Using Friction Surfacing Technique
by Hemlata Jangid, Nirmal K. Singh and Amlan Kar
Processes 2025, 13(3), 637; https://doi.org/10.3390/pr13030637 - 24 Feb 2025
Abstract
Friction surfacing (FS) is increasingly recognized as an advanced technique for coating similar and dissimilar materials, enabling superior joint quality through plastic deformation and grain refinement. This study investigates the deposition of AA6082-T651 alloy on a medium-carbon steel EN14B substrate using FS, with [...] Read more.
Friction surfacing (FS) is increasingly recognized as an advanced technique for coating similar and dissimilar materials, enabling superior joint quality through plastic deformation and grain refinement. This study investigates the deposition of AA6082-T651 alloy on a medium-carbon steel EN14B substrate using FS, with process parameters optimized, and the effect of axial load, rotational speed, and traverse speed on coating integrity. The optimal sample was subjected to heat treatment (HT) at 550 °C for 24, 36, and 48 h to further enhance mechanical properties. Comprehensive microstructural and mechanical analyses were performed on both heat-treated and non-heat-treated samples using optical microscopy (OM), field emission scanning electron microscopy (FESEM) with energy-dispersive X-ray spectroscopy (EDS), X-ray diffraction (XRD), microhardness testing, and micro-tensile techniques. The optimized sample was processed with a 6 kN axial load, a rotational speed of 2700 rpm, and a traverse speed of 400 mm/min, and demonstrated superior bond quality and enhanced mechanical properties. The highest interfacial hardness values, 138 HV0.1 were achieved for the sample annealed for 48 h, under an axial load of 6 kN. Annealing for 48 h significantly improved atomic bonding at the aluminum–steel interface, confirmed by the formation of Fe3Al intermetallic compounds detected via FESEM-EDS and XRD. These compounds were the primary reason for the enhancement in the mechanical properties of the FS deposit. Furthermore, the interrelationship between process and thermal parameters revealed that a peak temperature of 422 °C, heat input of 1.1 kJ/mm, and an axial load of 6 kN are critical for achieving optimal mechanical interlocking and superior coating quality. The findings highlight that optimized FS parameters and post-heat treatment are critical in achieving high-quality, durable coatings, with improved interfacial bonding and hardness, making the process suitable for structural applications. Full article
(This article belongs to the Special Issue Advances and Implementation of Welding and Additive Manufacturing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Schematic diagram of the friction surfacing with process parameters, (<b>b</b>) FSW machine setup, and (<b>c</b>) consumable tool after FS deposition.</p>
Full article ">Figure 2
<p>FSed deposition of AA6082 on EN14B at different processing conditions. Axial load/rotation speed/traverse speed are shown in each image.</p>
Full article ">Figure 3
<p>Influence of axial load and rotational speed on the bond quality, showing a window of good and no bonding.</p>
Full article ">Figure 4
<p><b>(a)</b> Al–steel coating transition area for Sample S9 and (<b>b</b>) magnified view of the cross-section of Sample S9.</p>
Full article ">Figure 5
<p>Optical micrograph associated with (<b>a</b>) NHT and (<b>b</b>) HT aluminum deposits during FS of the optimized sample (S9).</p>
Full article ">Figure 6
<p>FSed coating of NHT samples obtained at (<b>a</b>) 5 KN (S5), (<b>b</b>) 6 KN (S9), and (<b>c</b>) magnified FESEM image of optimized coating sample.</p>
Full article ">Figure 7
<p>FESEM images of heat-treated (HT) samples for (<b>a</b>) 24 h, (<b>b</b>) 36 h, (<b>c</b>) 48 h, and (<b>d</b>) a magnified image of HT for 48 h.</p>
Full article ">Figure 8
<p>(<b>a</b>) Al–steel interface at Spectrum 2, and (<b>b</b>–<b>d</b>) depict the EDS information of Samples S9 at 24 h, 36 h, and 48 h at the interface.</p>
Full article ">Figure 9
<p>(<b>a</b>) Al–steel interface at Spectrum 3, and (<b>b</b>–<b>d</b>) depict the EDS information of Sample S9 at 24 h, 36 h, and 48 h at the interface.</p>
Full article ">Figure 10
<p>XRD pattern analysis of EN14B steel coated with aluminum AA6082 using the FSed method.</p>
Full article ">Figure 11
<p>Microhardness of AA6082-coated FSed samples across the interface.</p>
Full article ">Figure 12
<p>Influence of process parameters on the microhardness of the FSed coating.</p>
Full article ">Figure 13
<p>Influence of process parameters on ultimate tensile strength (UTS) of the FSed coating.</p>
Full article ">Figure 14
<p>Tensile testing of AA6082-coated EN14B steel specimen using FS process.</p>
Full article ">Figure 15
<p>Variation of (<b>a</b>) Load and, (<b>b</b>) torque profile vs. processing time obtained from FSW machine data.</p>
Full article ">
12 pages, 3289 KiB  
Article
The Mechanical Performance Enhancement of the CrN/TiAlCN Coating on GCr15 Bearing Steel by Controlling the Nitrogen Flow Rate in the Transition Layer
by Yuchuan Cheng, Junxiang Li, Fang Liu, Hongjun Li and Nu Yan
Coatings 2025, 15(3), 254; https://doi.org/10.3390/coatings15030254 - 20 Feb 2025
Abstract
The main focus of this work is the successful deposition of hard and wear-resistant TiAlCN coating on the surface of GCr15 bearing steel by means of magnetron sputtering technology. The phase composition of the chromium nitride transition layer was monitored by precisely controlling [...] Read more.
The main focus of this work is the successful deposition of hard and wear-resistant TiAlCN coating on the surface of GCr15 bearing steel by means of magnetron sputtering technology. The phase composition of the chromium nitride transition layer was monitored by precisely controlling the nitrogen (N2) flow rate to strengthen the bonding between the TiAlCN coating and the GCr15 bearing steel surface. It was found that coating performance reached the optimal state at a N2 flow rate of 40 sccm, yielding a hardness of 23.3 GPa, a friction coefficient of only 0.27, and a wear rate of 0.19 × 10−8 mm3/N·m. Full article
Show Figures

Figure 1

Figure 1
<p>XRD patterns of CrN coatings deposited at different N<sub>2</sub> flow rates.</p>
Full article ">Figure 2
<p>XRD patterns of multilayered CrN/TiAlCN coatings deposited at different N<sub>2</sub> flow rates.</p>
Full article ">Figure 3
<p>(<b>a</b>–<b>e</b>) SEM images and (<b>a<sub>1</sub></b>–<b>e<sub>1</sub></b>) corresponding EDS elemental mapping of CrN coatings deposited at different N<sub>2</sub> flow rates: (<b>a</b>,<b>a<sub>1</sub></b>) 100 sccm, (<b>b</b>,<b>b<sub>1</sub></b>) 20 sccm, (<b>c</b>,<b>c<sub>1</sub></b>) 30 sccm, (<b>d</b>,<b>d<sub>1</sub></b>) 40 sccm, and (<b>e</b>,<b>e<sub>1</sub></b>) 50 sccm.</p>
Full article ">Figure 4
<p>(<b>a</b>) SEM images and (<b>b</b>) corresponding EDS elemental mapping of TiAlCN coatings deposited.</p>
Full article ">Figure 5
<p>The thickness of the CrN coating (<b>a</b>) and TiAlCN coating (<b>b</b>).</p>
Full article ">Figure 6
<p>Hardness and elastic modulus of (<b>a</b>) CrN and (<b>b</b>) CrN/TiAlCN coatings at different N<sub>2</sub> flow rates.</p>
Full article ">Figure 7
<p>(<b>a</b>) Scratch test results for the TiAlCN coating. (<b>b</b>–<b>f</b>) Scratch test results for the multilayered CrN/TiAlCN coating.</p>
Full article ">Figure 8
<p>(<b>a</b>) Frictional force and (<b>b</b>) friction coefficient of multilayered CrN/TiAlCN coatings deposited at different N<sub>2</sub> flow rates.</p>
Full article ">Figure 9
<p>Wear rates of multilayered CrN/TiAlCN coatings deposited at different N<sub>2</sub> flow rates.</p>
Full article ">
15 pages, 2959 KiB  
Article
Structural and Textural Properties of Al/Cu- and Al/Zn-Pillared Clays for Ethanol Conversion
by Lamara M. dos Santos, Felipe F. Barbosa, Lindiane Bieseki, Tiago P. Braga and Sibele B. C. Pergher
Crystals 2025, 15(3), 203; https://doi.org/10.3390/cryst15030203 - 20 Feb 2025
Abstract
A montmorillonite sample was pillared using mixed solutions of Al/Cu and Al/Zn as the pillaring agent. Al/Cu- and Al/Zn-pillared clays were applied in an ethanol conversion reaction. The catalysts prepared from montmorillonite were characterized using X-ray diffraction (XRD), X-ray fluorescence (XRF), 27Al nuclear [...] Read more.
A montmorillonite sample was pillared using mixed solutions of Al/Cu and Al/Zn as the pillaring agent. Al/Cu- and Al/Zn-pillared clays were applied in an ethanol conversion reaction. The catalysts prepared from montmorillonite were characterized using X-ray diffraction (XRD), X-ray fluorescence (XRF), 27Al nuclear magnetic resonance spectroscopy (MAS-NMR), and textural analysis by adsorption of N2. The synthesized materials showed basal spacing values around 1.6 nm for the Cu samples and 1.8 nm for the Zn samples. With the pillarization, textural properties such as specific area and micropore volume were optimized, and all samples showed an increase in micropore volume as well as a narrower pore distribution in the range of 0.77 nm. The insertion of 10% Zn in the pillaring solution produced a material with a greater amount of Al in the pentacoordinate position and also presented better results of conversion (80%) and selectivity to ethylene (81%) in the ethanol dehydration reaction. Full article
(This article belongs to the Section Inorganic Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>XRD patterns of montmorillonite and intercalated material with the ion of Keggin Al<sub>13</sub> (Al100), and with 10% Cu (Al/Cu10), 10% Zn (Al/Zn10), and 20% Zn (Al/Zn20). The crystalline phases are indicated: e = smectite, q = quartz, and cb = cristobalite. The symbol * indicates collapsed blades [<a href="#B3-crystals-15-00203" class="html-bibr">3</a>].</p>
Full article ">Figure 2
<p>XRD patterns of calcined montmorillonite and pillared clays (after calcination) with Keggin ion Al<sub>13</sub> (Al 100C), and with 10% Cu (Al/Cu 10C), 10% Zn (Al/Zn 10C), and 20% Zn (Al/Zn 20C). The crystalline phases are indicated: e = smectite, q = quartz, and cb = cristobalite. The symbol * indicates collapsed blades [<a href="#B3-crystals-15-00203" class="html-bibr">3</a>].</p>
Full article ">Figure 3
<p>(<b>a</b>) N<sub>2</sub> adsorption isotherms of montmorillonite and pillared clay. Closed symbols (▲) represent adsorption data, and open symbols (○) represent desorption data. (<b>b</b>) Micropore size distribution for montmorillonite and pillared samples using the Horvath and Kawazoe method applied to N<sub>2</sub> adsorption.</p>
Full article ">Figure 4
<p>Spectra of <sup>27</sup>Al (MAS) NMR for the following samples: (<b>a</b>) Montmorillonite, (<b>b</b>) pillared with Al, (<b>c</b>) containing 10% Cu, (<b>d</b>) containing 10% Zn, and (<b>e</b>) containing 20% Zn.</p>
Full article ">Figure 5
<p>Conversion and selectivity of the reaction with ethanol for the following clays: (<b>a</b>) montmorillonite, (<b>b</b>) Al100, (<b>c</b>) Al/Cu 10C, (<b>d</b>) Al/Zn 10C, and (<b>e</b>) Al/Zn20C.</p>
Full article ">Figure 6
<p>Proposed mechanism from reactions with ethanol by routs I, II and III.</p>
Full article ">
11 pages, 2098 KiB  
Article
The Facile Synthesis of Exogenous Lewis-Base-Free Amidoalanes: A Structural Comparison
by Jake Hemsworth, Andrej Vinogradov, William Lewis, Simon Woodward and Darren Willcox
Molecules 2025, 30(5), 986; https://doi.org/10.3390/molecules30050986 - 20 Feb 2025
Abstract
A simple one-pot reaction of LiAlH4, AlCl3, and a secondary amine HNR2 [R = Et, iPr, iBu, cyclo-C6H11, (CH2)4, and (CH2)5] in hydrocarbon solvents [...] Read more.
A simple one-pot reaction of LiAlH4, AlCl3, and a secondary amine HNR2 [R = Et, iPr, iBu, cyclo-C6H11, (CH2)4, and (CH2)5] in hydrocarbon solvents results in the formation of exogenous Lewis-base-free amidoalanes [H2Al(NR2)]n (n = 2 or 3) as crystalline solids (35–85% yield). In the solid state (seven X-ray structures), all the amidoalanes exist as dimers, with the exception of the pyrrolidine-derived alane which exists as a trimer. As solids, these amidoalanes exhibit significant kinetic stability towards oxygen/moisture allowing the brief (ca. 5 min.) handling of [H2Al(NiPr2)]2 in air. Full article
(This article belongs to the Section Organometallic Chemistry)
Show Figures

Figure 1

Figure 1
<p>Reported organoaluminium with enhanced air stability [<a href="#B1-molecules-30-00986" class="html-bibr">1</a>,<a href="#B2-molecules-30-00986" class="html-bibr">2</a>,<a href="#B3-molecules-30-00986" class="html-bibr">3</a>,<a href="#B4-molecules-30-00986" class="html-bibr">4</a>,<a href="#B5-molecules-30-00986" class="html-bibr">5</a>].</p>
Full article ">Figure 2
<p>Solid-state structures of amidoalanes: (<b>a</b>) <b>1a</b> [H<sub>2</sub>Al(NEt<sub>2</sub>)]<sub>2</sub>; (<b>b</b>) <b>1b</b> [H<sub>2</sub>Al (NPr<span class="html-italic"><sup>i</sup></span><sub>2</sub>)]<sub>2</sub>; (<b>c</b>) <b>1d</b> [H<sub>2</sub>Al(NBu<span class="html-italic"><sup>i</sup></span>)]<sub>2</sub>; (<b>d</b>) <b>1e</b> [H<sub>2</sub>Al (NCy<sub>2</sub>)]<sub>2</sub>; (<b>e</b>) <b>1f</b> [H<sub>2</sub>Al(NC<sub>4</sub>H<sub>8</sub>)]<sub>3</sub>; (<b>f</b>) <b>2</b> [HAl(NBu<span class="html-italic"><sup>i</sup></span><sub>2</sub>)<sub>2</sub>]<sub>2</sub>. Displacement ellipsoids are drawn at the 50% probability level; C-bonded hydrogen atoms are omitted for clarity.</p>
Full article ">Figure 3
<p>Air stability tests of <b>1b</b>, <b>1c</b>, <b>1d</b>, <b>1f</b>, and <b>2</b>.</p>
Full article ">Scheme 1
<p>Synthesis of amidoalanes.</p>
Full article ">Scheme 2
<p>Synthesis of diamidoalane (<b>2</b>).</p>
Full article ">Scheme 3
<p>Preliminary reactivity studies using amidoalane <b>1b</b>.</p>
Full article ">
18 pages, 300 KiB  
Article
Oral Health Status and Factors Associated with Oral Health in Patients with Alzheimer’s Disease: A Matched Case-Control Observational Study
by Reza Aghasizadeh Sherbaf, George Michael Kaposvári, Katalin Nagy, Magdolna Pakáski, Márió Gajdács, Danica Matusovits and Zoltán Baráth
J. Clin. Med. 2025, 14(5), 1412; https://doi.org/10.3390/jcm14051412 - 20 Feb 2025
Abstract
Background: Alzheimer’s disease (AD) is a chronic neurodegenerative disease, ranking as the seventh leading cause of death in both sexes. There is increasing awareness of the role of chronic periodontal disease and severe tooth loss as a modifiable risk factor for developing AD. [...] Read more.
Background: Alzheimer’s disease (AD) is a chronic neurodegenerative disease, ranking as the seventh leading cause of death in both sexes. There is increasing awareness of the role of chronic periodontal disease and severe tooth loss as a modifiable risk factor for developing AD. The aim of the present observational study was to assess AD patients with non-affected healthy controls in the context of their dental and periodontal health outcomes; additionally, the potential impact of anamnestic factors and lifestyle habits on oral health outcomes was also studied. Methods: A total of n = 41 AD patients receiving treatment at the Department of Psychiatry, University of Szeged, were compared with n = 41 age- and gender-matched controls from individuals seeking dental treatment and from retirement homes (mean age was 83.32 ± 7.82 years). Dental and periodontal status indices were assessed according to World Health Organization (WHO) criteria. Results: Overall, 51.2%, 68.3%, and 87.8% of AD patients received mood stabilizers, drugs for their non-cognitive symptoms and cognitive symptoms, respectively. Severe tooth loss was observed in 43.9% of AD patients and 56.1% of controls, respectively. There were no significant differences among AD patients and controls regarding the dental status indices studied (p > 0.05 for all indicators). AD patients had significantly higher plaque indices (%) (59.06 ± 15.45 vs. 41.35 ± 7.97; p < 0.001), bleeding on probing (BOP%) (62.65 ± 12.00 vs. 40.12 ± 10.86; p < 0.001), pocket depth [PD] (2.63 ± 0.56 vs. 2.29 ± 0.13; p = 0.002) and attachment loss [AL] (2.85 ± 0.79 vs. 2.39 ± 0.41; p = 0.026) values, compared to controls. Smoking (vs. non-smokers; 56.28 ± 12.36 vs. 51.40 ± 13.23, p = 0.038) and consumption of alcohol (vs. non-drinkers; 58.68 ± 9.86 vs. 54.78 ± 14.86, p = 0.040) were associated with higher plaque indices [%], while no similar effects were shown for dental status parameters (p > 0.05). In contrast, coffee intake and vitamin supplement use had no significant effect on dental or periodontal status parameters (p > 0.05 in all cases). Conclusions: The results of our study underscore the substantial treatment needs of AD patients, calling for heightened awareness among dental healthcare professionals. Full article
(This article belongs to the Special Issue Oral Hygiene: Updates and Clinical Progress)
16 pages, 5171 KiB  
Article
Effect of the Catalyst Support on the NOX Formation During Combustion of NH3 SOFC Off-Gas
by Tobias Weissenberger, Ralf Zapf, Helmut Pennemann and Gunther Kolb
Catalysts 2025, 15(3), 196; https://doi.org/10.3390/catal15030196 - 20 Feb 2025
Abstract
Ammonia has attracted much interest as a potential green and renewable hydrogen carrier or energy vector. Compared to hydrogen, ammonia offers several advantages. For example, ammonia has a significantly higher energy density and can be liquefied at room temperature at a moderate pressure [...] Read more.
Ammonia has attracted much interest as a potential green and renewable hydrogen carrier or energy vector. Compared to hydrogen, ammonia offers several advantages. For example, ammonia has a significantly higher energy density and can be liquefied at room temperature at a moderate pressure of 8 bars. While ammonia can be cracked to supply hydrogen, it is also possible to convert it directly into high-temperature solid oxide fuel cells (SOFCs) to generate electricity. The Ship-FC project aims to install an ammonia-fed 2MW SOFC system on board the vessel Viking energy to demonstrate the feasibility of zero CO2 emission shipping. For this NH3 SOFC system, a catalytic afterburner is required to remove the hydrogen and ammonia present in the SOFC off-gas and to recover heat. The current study analysed the effects of different catalyst supports, with a focus on NOX formation through the combustion of an SOFC off-gas surrogate. The study investigated the performance of catalysts based on the active metals, platinum and iridium, as well as the catalyst supports, Al2O3, SiO2, and TiO2. The results were correlated with catalyst characterisation data and ammonia TPD results. The investigations showed that the formation of NOX was clearly affected by the nature of the catalyst support. The highest selectivity towards NOX was observed for Al2O3, followed by SiO2, and the lowest selectivity was observed for TiO2. This trend was evident for the supported platinum and iridium catalysts and for the samples exclusively containing the support. The trend for N2O formation was opposite to that of NOX formation (TiO2 > SiO2 > Al2O3) in both the presence and absence of platinum or iridium. Full article
Show Figures

Figure 1

Figure 1
<p>Nitrogen sorption isotherms of platinum-based catalysts.</p>
Full article ">Figure 2
<p>Nitrogen sorption isotherms of the iridium-based catalysts.</p>
Full article ">Figure 3
<p>TEM micrographs of platinum-based catalysts.</p>
Full article ">Figure 4
<p>TEM micrographs of iridium-based catalysts.</p>
Full article ">Figure 5
<p>Ammonia temperature-programmed desorption (NH<sub>3</sub>-TPD) results for different supports (heating ramp 10 K min<sup>−1</sup>).</p>
Full article ">Figure 6
<p>(<b>a</b>) Hydrogen conversion and (<b>b</b>) ammonia conversion vs. reaction temperature for 5 wt.% platinum on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 7
<p>(<b>a</b>) NO<sub>X</sub> concentration and (<b>b</b>) NO<sub>X</sub> selectivities vs. reaction temperature for 5 wt.% platinum on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 8
<p>(<b>a</b>) N<sub>2</sub>O concentration and (<b>b</b>) N<sub>2</sub>O selectivities over reaction temperature for 5 wt.% platinum on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 9
<p>(<b>a</b>) Hydrogen and (<b>b</b>) ammonia conversion vs. reaction temperature for 5 wt.% iridium on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 10
<p>(<b>a</b>) NO<sub>X</sub> concentration and (<b>b</b>) NO<sub>X</sub> selectivities vs. reaction temperature for 5 wt.% iridium on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 11
<p>(<b>a</b>) N<sub>2</sub>O concentration and (<b>b</b>) N<sub>2</sub>O selectivities over reaction temperature for 5 wt.% iridium on different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 12
<p>(<b>a</b>) NH<sub>3</sub> Conversion over temperature for different supports, (<b>b</b>) NO<sub>X</sub> concentrations over temperature for different supports, WHSV = 600 L/g h.</p>
Full article ">Figure 13
<p>Measured NO<sub>X</sub> concentrations over different catalysts at 700 °C.</p>
Full article ">
17 pages, 1770 KiB  
Article
Revisiting the Mechanistic Pathway of Gas-Phase Reactions in InN MOVPE Through DFT Calculations
by Xiaokun He, Nan Xu, Yuan Xue, Hong Zhang, Ran Zuo and Qian Xu
Molecules 2025, 30(4), 971; https://doi.org/10.3390/molecules30040971 - 19 Feb 2025
Abstract
III-nitrides are crucial materials for solar flow batteries due to their versatile properties. In contrast to the well-studied MOVPE reaction mechanism for AlN and GaN, few works report gas-phase mechanistic studies on the growth of InN. To better understand the reaction thermodynamics, this [...] Read more.
III-nitrides are crucial materials for solar flow batteries due to their versatile properties. In contrast to the well-studied MOVPE reaction mechanism for AlN and GaN, few works report gas-phase mechanistic studies on the growth of InN. To better understand the reaction thermodynamics, this work revisited the gas-phase reactions involved in metal–organic vapor-phase epitaxy (abbreviated as MOVPE) growth of InN. Utilizing the M06-2X function in conjunction with Pople’s triple-ζ split-valence basis set with polarization functions, this work recharacterized all stationary points reported in previous literature and compared the differences between the structures and reaction energies. For the reaction pathways which do not include a transition state, rigorous constrained geometry optimizations were utilized to scan the PES connecting the reactants and products in adduct formation and XMIn (M, D, T) pyrolysis, confirming that there are no TSs in these pathways, which is in agreement with the previous findings. A comprehensive bonding analysis indicates that in TMIn:NH3, the In-N demonstrates strong coordinate bond characteristics, whereas in DMIn:NH3 and MMIn:NH3, the interactions between the Lewis acid and base fragments lean toward electrostatic attraction. Additionally, the NBO computations show that the H radical can facilitate the migration of electrons that are originally distributed between the In-C bonds in XMIn. Based on this finding, novel reaction pathways were also investigated. When the H radical approaches MMInNH2, MMIn:NH3 rather than MMInHNH2 will generate and this is followed by the elimination of CH4 via two parallel paths. Considering the abundance of H2 in the environment, this work also examines the reactions between H2 and XMIn. The Mulliken charge distributions indicated that intermolecular electron transfer mainly occurs between the In atom and N atom whiling forming (DMInNH2)2, whereas it predominately occurs between the In atom and the N atom intramolecularly when generating (DMInNH2)3. Full article
(This article belongs to the Section Physical Chemistry)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Two parallel paths with the elimination of CH<sub>4</sub> from MMIn:NH<sub>3</sub> and corresponding molecular structures.</p>
Full article ">Figure 2
<p>The relaxed scan for adduct formation (A1–A1b) and pyrolysis reaction (P4–P4b). [Annotation 1] The PES was explored by constrained geometry optimization, and connects the dissociated In(CH<sub>3</sub>)<sub>x−1</sub> and CH<sub>3</sub> or In(CH<sub>3</sub>)<sub>x</sub> and NH<sub>3</sub>. [Annotation 2] Relative Energy refers to the electron energy difference between the scan points and the 1st scan point (i.e., reactants).</p>
Full article ">Figure 3
<p>The ESP map of TMIn and NH<sub>3</sub>.</p>
Full article ">Figure 4
<p>The HOMO and LUMO and the associated <span class="html-italic">E</span><sub>gap</sub> of TS in reactions A1, A1a and A1b.</p>
Full article ">Figure 5
<p>The HOMO and LUMO and the <span class="html-italic">E</span><sub>gap</sub> (in eV) of TMIn, DMIn and MMIn.</p>
Full article ">Figure 6
<p>The critical bond lengths and atom distances (in Å) along with bond angles (in <sup>o</sup>) in the fully optimized TS of R9.</p>
Full article ">Figure 7
<p>The ESP map of DMInNH<sub>2</sub>.</p>
Full article ">
16 pages, 288 KiB  
Article
Donsker-Type Theorem for Numerical Schemes of Backward Stochastic Differential Equations
by Yi Guo and Naiqi Liu
Mathematics 2025, 13(4), 684; https://doi.org/10.3390/math13040684 - 19 Feb 2025
Abstract
This article studies the theoretical properties of the numerical scheme for backward stochastic differential equations, extending the relevant results of Briand et al. with more general assumptions. To be more precise, the Brown motion will be approximated using the sum of a sequence [...] Read more.
This article studies the theoretical properties of the numerical scheme for backward stochastic differential equations, extending the relevant results of Briand et al. with more general assumptions. To be more precise, the Brown motion will be approximated using the sum of a sequence of martingale differences or a sequence of i.i.d. Gaussian variables instead of the i.i.d. Bernoulli sequence. We cope with an adaptation problem of Yn by defining a new process Y^n; then, we can obtain the Donsker-type theorem for numerical solutions using a similar method to Briand et al. Full article
56 pages, 16932 KiB  
Review
Study of the Influence of Nanoparticle Reinforcement on the Mechanical and Tribological Performance of Aluminum Matrix Composites—A Review
by Varun Singhal, Daksh Shelly, Abhishek Saxena, Rahul Gupta, Vipin Kumar Verma and Appurva Jain
Lubricants 2025, 13(2), 93; https://doi.org/10.3390/lubricants13020093 - 19 Feb 2025
Abstract
This study investigates the influence of nano-sized reinforcements on aluminum matrix composites’ mechanical and tribological properties. Microstructural analysis revealed that introducing nanoparticles led to grain refinement, reducing the grain size from 129.7 μm to 41.3 μm with 2 wt.% TiO2 addition. Furthermore, [...] Read more.
This study investigates the influence of nano-sized reinforcements on aluminum matrix composites’ mechanical and tribological properties. Microstructural analysis revealed that introducing nanoparticles led to grain refinement, reducing the grain size from 129.7 μm to 41.3 μm with 2 wt.% TiO2 addition. Furthermore, ultrasonic-assisted squeeze casting of AA6061 composites reinforced with TiO2 and Al2O3 resulted in a 52% decrease in grain size, demonstrating nano-reinforcements’ effectiveness in refining the matrix structure. Despite these advantages, the high surface energy of nanoparticles causes agglomeration, which can undermine composite performance. However, ultrasonic-assisted stir casting reduced agglomeration by approximately 80% compared to conventional stir casting, and cold isostatic pressing improved dispersion uniformity by 27%. The incorporation of nano-reinforcements such as SiC, Al2O3, and TiC significantly enhanced the material properties, with hardness increasing by ~30% and ultimate tensile strength improving by ~80% compared to pure Al. The hardness of nano-reinforced composites substantially rose from 83 HV (pure Al) to 117 HV with 1.0 vol.% CNT reinforcement. Additionally, TiC-reinforced AA7075 composites improved hardness from 94.41 HB to 277.55 HB after 10 h of milling, indicating a nearly threefold increase. The wear resistance of Al-Si alloys was notably improved, with wear rates reduced by up to 52%, while the coefficient of friction decreased by 20–40% with the incorporation of graphene and CNT reinforcements. These findings highlight the potential of nano-reinforcements in significantly improving the mechanical and tribological performance of n-AMCs, making them suitable for high-performance applications in aerospace, automotive, and structural industries. Full article
Show Figures

Figure 1

Figure 1
<p>Classification of n-AMC manufacturing processes.</p>
Full article ">Figure 2
<p>(<b>a</b>) Year-wise distribution of published documents from 2021 to 2024, showing an increasing trend with a peak in 2023. (<b>b</b>) Document type and subject category analysis, where articles make up the majority (76%), followed by conference papers (13%), book chapters (4%), conference reviews (4%), and reviews (3%). (Source of Scopus).</p>
Full article ">Figure 3
<p>Classification of reinforcements in fabrication of n-AMCs.</p>
Full article ">Figure 4
<p>EBSD maps of grain structure: (<b>a</b>) Al, (<b>b</b>) Al-2.5%CNT-1%γ-Al<sub>2</sub>O<sub>3</sub>, (<b>c</b>) Al-7.5%CNT-1%γ-Al<sub>2</sub>O<sub>3</sub>. Dark areas indicate confidence index &lt; 0.1 [<a href="#B95-lubricants-13-00093" class="html-bibr">95</a>].</p>
Full article ">Figure 5
<p>(<b>a</b>) The TEM image of the GO-CNT/Al composite shows dislocation entanglement at grain boundaries. (<b>b</b>) TEM image highlighting Al<sub>4</sub>C<sub>3</sub> nanorods (red ellipses). (<b>c</b>) HRTEM image of the Al/Al<sub>4</sub>C<sub>3</sub> interface with (003) lattice spacing of 0.83 nm. (<b>d</b>) Lattice details. (<b>e</b>) Schematic of Al<sub>4</sub>C<sub>3</sub> formation across sintering, hot extrusion, and extrusion stages [<a href="#B101-lubricants-13-00093" class="html-bibr">101</a>].</p>
Full article ">Figure 6
<p>Microstructural analysis and grain size distribution: (<b>a</b>) TEM image of CNT aligned along the rolling direction. (<b>b</b>) HRTEM showing stacking faults and the Al-CNT interface. (<b>c</b>) Magnified view of the stacking fault region. (<b>d</b>) SAED pattern along the [011] axis of Al. (<b>e</b>) EBSD map of grain orientation with a color legend inset. (<b>f</b>) Grain size distribution histogram, averaging 1.53 ± 0.16 μm [<a href="#B104-lubricants-13-00093" class="html-bibr">104</a>].</p>
Full article ">Figure 7
<p>Microstructural characterization of Al-(TiB<sub>2</sub>/Cu) composites via SEM and elemental mapping. (<b>a</b>,<b>b</b>) Low-magnification SEM images showing micropores, agglomerations, and TiB<sub>2</sub>/Cu phase distribution. (<b>c</b>–<b>f</b>) Higher magnification SEM images detailing TiB<sub>2</sub>/Cu phase morphology in the α-Al matrix. (<b>g</b>,<b>h</b>) Elemental mapping of Al, Cu, and Ti, highlighting their spatial distribution [<a href="#B106-lubricants-13-00093" class="html-bibr">106</a>].</p>
Full article ">Figure 8
<p>HRTEM images of Al composites processed by SPS and LPBF. (<b>a</b>) SPS sample showing PPB and SiC phases. (<b>b</b>) HRTEM of region b with 0.254 nm lattice spacing for SiC (012) plane. (<b>c</b>) The LPBF sample shows Si particle distribution. (<b>d</b>) HRTEM of region d with 0.311 nm lattice spacing for Si (111) plane. (<b>e</b>) LPBF sample with Al<sub>4</sub>C<sub>3</sub> phase. (<b>f</b>) FFT of region f, showing (011) and (003) planes of Al<sub>4</sub>C<sub>3</sub> [<a href="#B110-lubricants-13-00093" class="html-bibr">110</a>].</p>
Full article ">Figure 9
<p>SEM images and EDS mapping of an Al-Ti-O-Zn composite. (<b>a</b>) SEM image showing pores and clusters. (<b>b</b>) EDS mapping highlighting Al, Ti, O, and Zn distribution, with Al dominant. (<b>c</b>) Detailed EDS maps of Al (red), Ti (black), O (white), and Zn (blue). (<b>d</b>) SEM image of an agglomeration. Insets show corresponding EDS spectra [<a href="#B111-lubricants-13-00093" class="html-bibr">111</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) SEM micrograph of Al-4TiCn composite showing uniformly dispersed nTiC particles (yellow dashed circles) within the Al matrix. (<b>b</b>) EDX spectrum confirms the elemental composition, with Al as the major element and trace amounts of Ti, Si, Fe, Ag, and O [<a href="#B113-lubricants-13-00093" class="html-bibr">113</a>].</p>
Full article ">Figure 11
<p>Microstructural polarized images of (<b>a</b>) H1 (1200 rpm rotational speed), (<b>b</b>) H2 (1300 rpm rotational speed), (<b>c</b>) H3 (1400 rpm rotational speed), (<b>d</b>) H4 (1500 rpm rotational speed), and (<b>e</b>) reference (AA6061 base sample) (1200 rpm rotational speed) [<a href="#B117-lubricants-13-00093" class="html-bibr">117</a>].</p>
Full article ">Figure 12
<p>Microstructural analysis of samples: (<b>a</b>–<b>c</b>) EBSD grain orientation maps showing equiaxed grains, (<b>d</b>–<b>f</b>) KAM maps highlighting strain distribution. Average grain sizes: (<b>d</b>) 729 nm, (<b>e</b>) 954 nm, and (<b>f</b>) 649 nm. KAM values: (<b>d</b>) 0.31°, (<b>e</b>) 0.21°, and (<b>f</b>) 0.34°, indicating varying local strain levels [<a href="#B118-lubricants-13-00093" class="html-bibr">118</a>].</p>
Full article ">Figure 13
<p>Microstructural analysis showing the dispersion of B<sub>4</sub>C reinforcement particles in the composite. (<b>a</b>) Uniformly distributed fine B<sub>4</sub>C particles, (<b>b</b>) agglomerated B<sub>4</sub>C particles, and (<b>c</b>) clear particle boundaries and clustering of B<sub>4</sub>C [<a href="#B125-lubricants-13-00093" class="html-bibr">125</a>].</p>
Full article ">Figure 14
<p>Microstructure of hot-extruded 7085Al and (ZrB<sub>2</sub> + Al<sub>2</sub>O<sub>3</sub>)/7085Al: (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>) 3D SEM images of the base metal, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) optical micrographs along the extrusion direction, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) grain size distribution statistics [<a href="#B139-lubricants-13-00093" class="html-bibr">139</a>].</p>
Full article ">Figure 15
<p>(<b>a</b>) Base 7N01 alloy shows coarse, heterogeneous grains. (<b>b</b>) 7N01/ZrB<sub>2</sub> composite exhibits grain refinement. (<b>c</b>) 7N01/Se composite shows Se-induced refinement. (<b>d</b>) 7N01/ZrB<sub>2</sub>/Se composite achieves the most refined, uniform structure. (<b>e</b>) Grain size distribution graph shows a reduction from 182.6 μm (7N01) to 44.3 μm (7N01/ZrB<sub>2</sub>), 33.7 μm (7N01/Se), and 24.2 μm (7N01/ZrB<sub>2</sub>/Se) [<a href="#B128-lubricants-13-00093" class="html-bibr">128</a>].</p>
Full article ">Figure 16
<p>(<b>a</b>) A low-magnification TEM image shows the overall structure of the material. (<b>b</b>) A high-resolution TEM image is displaying MWCNT, as indicated by the arrows. (<b>c</b>) HRTEM image of an MWCNT interface. The corresponding EDS elemental maps for Al (red), O (blue), C (green), and combined RGB mapping confirm the material composition and the clustering of oxygen near MWCNT [<a href="#B129-lubricants-13-00093" class="html-bibr">129</a>].</p>
Full article ">Figure 17
<p>(<b>a</b>) SEM image showing the microstructure with analysis boundaries (dashed lines). (<b>b</b>,<b>c</b>) High-magnification SEM images of MEG and FEG regions highlight morphological differences. (<b>d</b>–<b>f</b>) EBSD orientation maps showing grain distribution, with (<b>d</b>,<b>f</b>) representing FEGs and (<b>e</b>) MEGs. (<b>g</b>) Grain size distribution: MEGs (1.54 μm) and FEGs (0.71 μm). (<b>h</b>) Pole figures for {100}, {110}, and {111} planes indicate crystallographic texture in MEG and FEG regions [<a href="#B133-lubricants-13-00093" class="html-bibr">133</a>].</p>
Full article ">Figure 18
<p>(<b>a</b>) SEM micrograph of the sample. (<b>b</b>) EBSD map with grain boundaries and IPF color key. (<b>c</b>) Misorientation distribution: 48.91% HAGBs, 51.09% LAGBs, with a theoretical comparison. (<b>d</b>) KAM map showing local misorientation gradients. (<b>e</b>) Crystallographic texture analysis with pole figures ({100}, {110}, {111}) and ODF sections (φ<sub>2</sub> = 0°, 15°, 30°, 45°) [<a href="#B134-lubricants-13-00093" class="html-bibr">134</a>].</p>
Full article ">Figure 19
<p>(<b>a</b>) TEM image of TiAl phases and CNT in the Al matrix (CNTs@Ti/Al composite). (<b>b</b>) HRTEM of CNT (002) plane with 0.34 nm spacing. (<b>c</b>) TiAl (013) plane with 0.32 nm spacing and FFT inset. (<b>d</b>) Low-magnification TEM of TiAl near grain boundaries. (<b>e</b>) Magnified TiAl phase from (<b>d</b>). (<b>f</b>) FFT and lattice analysis confirming 0.41 nm spacing for TiAl (002) plane [<a href="#B135-lubricants-13-00093" class="html-bibr">135</a>].</p>
Full article ">Figure 20
<p>The figure shows a hardness comparison graph, with hardness values (in HV) plotted on the <span class="html-italic">y</span>-axis and sample labels H1 (1200 rpm rotational speed), H2 (1300 rpm rotational speed), H3 (1400 rpm rotational speed), H4 (1500 rpm rotational speed), and reference (AA6061 base sample) (1200 rpm rotational speed) on the <span class="html-italic">x</span>-axis.</p>
Full article ">Figure 21
<p>(<b>a</b>) Vickers hardness comparison for pure Al, Al-O, Al/MWCNT, and Al-O/MWCNT composites, showing hardness improvement with reinforcement. (<b>b</b>) Nanoindentation load–depth curves illustrating mechanical response differences. (<b>c</b>) Compressive stress–strain curves highlight strength improvement with reinforcement [<a href="#B129-lubricants-13-00093" class="html-bibr">129</a>].</p>
Full article ">Figure 22
<p>Contour maps illustrate a measured property’s spatial distribution across the sample surface. (<b>a</b>) Map of the property distribution for Sample X, showing relatively uniform regions with localized variations. (<b>b</b>) Map of the property distribution for Sample Y, highlighting higher intensity zones and increased non-uniformity compared to those of Sample X. The color scale indicates the magnitude of the property, with red representing higher values and blue indicating lower values [<a href="#B134-lubricants-13-00093" class="html-bibr">134</a>].</p>
Full article ">Figure 23
<p>(<b>a</b>,<b>b</b>) Stress–strain curves for 0°, 45°, and 90° extrusion angles under varying billet temperatures and ram speeds. (<b>c</b>) Effect of billet temperature on tensile stress (bars) and elongation (lines). (<b>d</b>) Influence of ram speed on tensile stress and elongation. (<b>c1</b>,<b>c2</b>,<b>d1</b>,<b>d2</b>) show linear trends with increasing temperature and speed [<a href="#B118-lubricants-13-00093" class="html-bibr">118</a>].</p>
Full article ">Figure 24
<p>(<b>a</b>) The stress–strain curves of the 7085Al alloy were reinforced with varying volume fractions (1, 3, and 5 vol.%) of ZrB<sub>2</sub> + Al<sub>2</sub>O<sub>3</sub> nanoparticle hybrid particles. (<b>b</b>) Engineering stress–strain curves of the same materials. The results illustrate the influence of increasing hybrid reinforcement volume fraction on the mechanical properties, particularly stress and strain behavior [<a href="#B139-lubricants-13-00093" class="html-bibr">139</a>].</p>
Full article ">Figure 25
<p>(<b>a</b>) Influence of ZrB<sub>2</sub> volume fraction (1, 3, and 5 vol.%); (<b>b</b>) comparison of 7N01Al alloy with and without SrSc addition; (<b>c</b>) combined consequence of ZrB<sub>2</sub> and Sc addition on the mechanical behavior. (<b>d</b>) ZrB<sub>2</sub> reinforcement volume fraction, (<b>e</b>) Sc addition, and (<b>f</b>) combined reinforcement and Sc effects [<a href="#B128-lubricants-13-00093" class="html-bibr">128</a>].</p>
Full article ">Figure 26
<p>(<b>a</b>) Stress–strain curves before and after 325 °C heat treatment (2 h) show improved mechanical behavior. (<b>b</b>) Strength contributions from σGB, σSS, and σP highlight increased YS and UTS with retained ductility [<a href="#B133-lubricants-13-00093" class="html-bibr">133</a>].</p>
Full article ">Figure 27
<p>The dimpled fracture in Al-CNT-γAl<sub>2</sub>O<sub>3</sub> AMCs: (<b>a</b>) As-fabricated microstructure, (<b>b</b>) nano-voids around γAl<sub>2</sub>O<sub>3</sub>, (<b>c</b>) CNT bridging that limits micro-void growth, (<b>d</b>) dimples with γAl<sub>2</sub>O<sub>3</sub> and ruptured CNT after coalescence [<a href="#B95-lubricants-13-00093" class="html-bibr">95</a>].</p>
Full article ">Figure 28
<p>(<b>a</b>) Overview of the fracture surface, highlighting the network structure. (<b>b</b>) Higher magnification reveals GO particles embedded within the matrix. (<b>c</b>) A detailed view shows the interface between GO and the matrix. (<b>d</b>) Schematic illustration of the strengthening mechanisms, including dislocation interaction with CNT and the formation of Orowan loops around GO particles, contributing to enhanced mechanical properties [<a href="#B101-lubricants-13-00093" class="html-bibr">101</a>].</p>
Full article ">Figure 29
<p>(<b>a</b>–<b>e</b>) SEM micrographs of fracture surfaces at different deformation levels, with the corresponding average dimple size labeled for each image (0.87 μm, 0.40 μm, 0.31 μm, 0.29 μm, and 0.27 μm). The progression shows dimple refinement and increased ductility with deformation. (<b>f</b>) High-magnification image highlighting GNPs distributed within the matrix, indicated by red arrows, illustrating their role in improving fracture toughness through energy dissipation and crack deflection mechanisms [<a href="#B103-lubricants-13-00093" class="html-bibr">103</a>].</p>
Full article ">Figure 30
<p>Tensile fracture surface morphology of porthole die extrusion specimens: (<b>a</b>) 480 °C/0.1 mm·s<sup>−1</sup>, (<b>b</b>) 540 °C/0.1 mm·s<sup>−1</sup>, and (<b>c</b>) 480 °C/0.5 mm·s<sup>−1</sup>. The top row shows low-magnification overviews, while the middle and bottom rows reveal detailed fracture features. (<b>a1</b>,<b>a2</b>,<b>a3</b>) Deep dimples indicate ductile failure at 480 °C/0.1 mm·s<sup>−1</sup>. (<b>b1</b>,<b>b2</b>,<b>b3</b>) Micro-voids and inclusions at 540 °C/0.1 mm·s<sup>−1</sup> suggest mixed-mode failure. (<b>c1</b>,<b>c2</b>,<b>c3</b>) Shallow dimples and micro-voids at 480 °C/0.5 mm·s<sup>−1</sup> indicate reduced ductility due to higher strain rates [<a href="#B118-lubricants-13-00093" class="html-bibr">118</a>].</p>
Full article ">Figure 31
<p>The figure presents SEM micrographs and EDS spectra of fracture surfaces. (<b>a</b>,<b>d</b>) Low-magnification views highlighting regions A and B. (<b>b</b>,<b>e</b>) Higher magnification images revealing voids and fracture features. (<b>c</b>,<b>f</b>) EDS spectra confirm Al, Zn, and Zr presence, indicating their role in fracture behavior and reinforcement [<a href="#B139-lubricants-13-00093" class="html-bibr">139</a>].</p>
Full article ">Figure 32
<p>TEM analysis of the CNT-Ti/Al AMC after tension: (<b>a</b>) Al<sub>3</sub>Ti trapping dislocations; (<b>b</b>) Al<sub>4</sub>C<sub>3</sub>, CNT, and nanoscale SFs; (<b>c</b>,<b>d</b>) high-magnification images of SFs; (<b>e</b>,<b>f</b>) GPA of the Al<sub>4</sub>C<sub>3</sub>-Al interface; (<b>g</b>,<b>h</b>) GPA of the Al<sub>3</sub>Ti-Al interface; (<b>i</b>) schematic of deformation mechanisms [<a href="#B135-lubricants-13-00093" class="html-bibr">135</a>].</p>
Full article ">Figure 33
<p>The variation in the COF with sliding distance for hybrid nanocomposites under different loads: (<b>a</b>) 10 N, (<b>b</b>) 30 N, and (<b>c</b>) 50 N. (<b>d</b>) summarizes the mean COF values for all n-AMCs at each load level (10 N, 30 N, and 50 N) [<a href="#B136-lubricants-13-00093" class="html-bibr">136</a>].</p>
Full article ">Figure 34
<p>(<b>a</b>) Significant grooves and debris accumulation indicate abrasive wear. (<b>b</b>) Severe wear with delamination and cracking suggests material instability. (<b>c</b>) Smoother wear tracks with minor debris imply improved wear resistance. (<b>d</b>) Deep grooves and detachment indicate adhesive wear under high stress. (<b>e</b>) Uniform wear tracks with minimal damage suggest enhanced tribological performance [<a href="#B117-lubricants-13-00093" class="html-bibr">117</a>].</p>
Full article ">Figure 35
<p>Schematic of tribological mechanism and wear progression in n-AMCs: (<b>a</b>) asperity interlocking with debris entrapment, (<b>b</b>) asperity distortion and debris accumulation, (<b>c</b>) tribomechanical layer (TML) formation, and (<b>d</b>) load-induced craters and debris expulsion [<a href="#B136-lubricants-13-00093" class="html-bibr">136</a>].</p>
Full article ">
19 pages, 4392 KiB  
Article
Fire Prevention and Extinguishing Characteristics of Al3+-CS/PAM-MBA Composite Dual-Network Gel
by Jianguo Wang, Yueyang Zhou, Yifan Zhao and Zhenzhen Zhang
Gels 2025, 11(2), 148; https://doi.org/10.3390/gels11020148 - 19 Feb 2025
Abstract
A physically and chemically cross-linked Al3+-CS/PAM-MBA dual-network gel with enhanced fire-suppression performance was prepared using chitosan (CS), acrylamide (AM), and N,N’-methylenebisacrylamide (MBA) as base materials. The first network was formed through the covalent cross-linking of polyacrylamide (PAM) with MBA, while the [...] Read more.
A physically and chemically cross-linked Al3+-CS/PAM-MBA dual-network gel with enhanced fire-suppression performance was prepared using chitosan (CS), acrylamide (AM), and N,N’-methylenebisacrylamide (MBA) as base materials. The first network was formed through the covalent cross-linking of polyacrylamide (PAM) with MBA, while the second network was established by crosslinking CS molecules with Al3+ ions. The optimal gel ratio was determined by evaluating its formation time and viscosity. The fire prevention and extinguishing performance of the gel was assessed through thermal stability analysis, temperature-programmed studies, infrared spectroscopy, thermal analysis, and fire-extinguishing experiments. The results indicated that the Al3+-CS/PAM-MBA dual-network gel exhibited excellent thermal stability and a strong self-ignition inhibition effect, effectively suppressing coal spontaneous combustion and oxidation. The gel achieved this by chemically inactivating coal molecules, disrupting the functional groups closely associated with coal–oxygen reactions and thereby hindering these reactions. Fire-extinguishing tests demonstrated that the gel restrained coal from spontaneous combustion. Upon application, the gel rapidly reduced the coal temperature, making re-ignition less likely. Full article
(This article belongs to the Special Issue Applications of Gels in Energy Materials and Devices)
Show Figures

Figure 1

Figure 1
<p>Effect of various factor levels on gelation time.</p>
Full article ">Figure 2
<p>Effect of various factor levels on gel viscosity.</p>
Full article ">Figure 3
<p>Variation curve of the water loss rate of two gels under a constant temperature.</p>
Full article ">Figure 4
<p>Water loss rate of the two gels under different temperature conditions.</p>
Full article ">Figure 5
<p>Curve of CO volume fraction as a function of temperature variation.</p>
Full article ">Figure 6
<p>Trend of change in inhibition rate.</p>
Full article ">Figure 7
<p>Curve of activation energy for each group with temperature variation.</p>
Full article ">Figure 8
<p>Infrared absorption spectra of different sample groups.</p>
Full article ">Figure 9
<p>TG-DSC curve of raw coal and composite dual-network gel-treated coal samples.</p>
Full article ">Figure 10
<p>Comparison of fire-extinguishing effects of different gel treatments.</p>
Full article ">Figure 11
<p>Diagram of the flame retardant mechanism of the composite dual-network gel.</p>
Full article ">Figure 12
<p>Preparation process for Al<sup>3+</sup>-CS/PAM-MBA composite dual-network gel.</p>
Full article ">Figure 13
<p>Diagram of the programmed heating device.</p>
Full article ">Figure 14
<p>Homemade fire-extinguishing test device.</p>
Full article ">
26 pages, 9359 KiB  
Article
Experimental and Numerical Analyses of the Influence of Al2O3 Nanoparticle Supplementation in Biodiesel (Water Hyacinth) Blends with Diesel on CI Engine Responses
by Ameer Hasan Hamzah, Abdulrazzak Akroot and Hasanain A. Abdul Wahhab
Appl. Sci. 2025, 15(4), 2204; https://doi.org/10.3390/app15042204 - 19 Feb 2025
Abstract
The current work includes experimental and numerical investigations into the effects of biodiesel (Eichhornia Crassipes) blends with different aluminum oxide nanoparticle concentrations on the combustion process in diesel engines. The experiments included measuring performance parameters and emissions tests while changing the engine speed [...] Read more.
The current work includes experimental and numerical investigations into the effects of biodiesel (Eichhornia Crassipes) blends with different aluminum oxide nanoparticle concentrations on the combustion process in diesel engines. The experiments included measuring performance parameters and emissions tests while changing the engine speed and increasing loads. IC Engine Fluent, a specialist computational tool included in the ANSYS software (R19.0 version), was used to simulate internal combustion engine dynamics and combustion processes. All investigations were carried out using biodiesel blends with three concentrations of Al2O3 nanoparticles: 50, 100, and 150 ppm. The tested samples are called D100, D80B20, D80B20N50, D80B20N100, and D80B20N150, accordingly. The combustion characteristics are improved due to the catalytic effect and higher surface area of nano additives. The results showed improvements in the combustion process as the result of the nanoparticles’ addition, which led to the higher peak cylinder pressure. The increases in the peak cylinder pressures for D80B20N50, D80B20N100, and D80B20N150 about D80B20 were 3%, 5%, and 8%, respectively, at a load of 200 Nm, while the simulation found that the maximum temperature for biodiesel blends diesel was higher than that for pure diesel; this was due to the higher hydrocarbon values of D80B20. Also, nano additives caused a decrease in temperatures in the combustion of biofuels. Finally, nano additives caused an enhancement of the emissions test results for all parameters when compared to pure diesel fuel and biofuel. Full article
(This article belongs to the Special Issue Clean Combustion Technologies and Renewable Fuels)
Show Figures

Figure 1

Figure 1
<p>Experimental setup visualization.</p>
Full article ">Figure 2
<p>Numerical analysis flowchart.</p>
Full article ">Figure 3
<p>Cylinder geometry of internal combustion engine.</p>
Full article ">Figure 4
<p>CI engine domain and mesh generated.</p>
Full article ">Figure 5
<p>Variation of the BSFC with load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 6
<p>Variation of the BTE with the load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 7
<p>Variation in the CO emissions with the load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 8
<p>Variation of the CO<sub>2</sub> emissions with the load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 9
<p>Variation of the HC emissions with the load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 10
<p>Variation of the NOx emissions with the load for all fuel blend samples: D100, D80B20, D80B20N50, D80B20N100, and D80B20N150.</p>
Full article ">Figure 11
<p>Contours of cylinder pressure variation with the crank angle of all tested fuels at a load of 200 Nm.</p>
Full article ">Figure 12
<p>Variation of cylinder pressure with the crank angle of all tested fuels at a load of 200 Nm.</p>
Full article ">Figure 13
<p>Peak cylinder pressure of the diesel and biodiesel blend with nano additives at different loads.</p>
Full article ">Figure 14
<p>Contours of cylinder temperature variation with the crank angle of B20N150 at a load of 200 Nm.</p>
Full article ">Figure 15
<p>Max. temperature of the diesel and biodiesel blend with nano additives at different loads.</p>
Full article ">Figure 16
<p>Contours of velocity distribution at different crank angles of D80B20N150 at a load of 200 Nm.</p>
Full article ">Figure 17
<p>The variation of the Numerical BTE with the engine load for fuel blend samples.</p>
Full article ">Figure 18
<p>Contours of CO<sub>2</sub> mass fraction of different fuel blends at a load of 200 Nm and speed of 1150 rpm.</p>
Full article ">Figure 19
<p>Contours of CO<sub>2</sub> mass fraction of different fuel blends at a load of 200 Nm and speed of 1400 rpm.</p>
Full article ">Figure 20
<p>Contours of CO<sub>2</sub> mass fraction of different fuel blends at a load of 200 Nm and speed of 1600 rpm.</p>
Full article ">Figure 21
<p>Contours of CO<sub>2</sub> mass fraction of different fuel blends at a load of 200 Nm and speed of 1800 rpm.</p>
Full article ">Figure 22
<p>Numerical results of CO<sub>2</sub> emissions with engine speed for all additives.</p>
Full article ">Figure 23
<p>Contours of NOx mass fraction of different fuel blends at a load of 200 Nm and speed of 1150 rpm.</p>
Full article ">Figure 24
<p>Contours of NOx mass fraction of different fuel blends at a load of 200 Nm and speed of 1400 rpm.</p>
Full article ">Figure 25
<p>Contours of NOx mass fraction of different fuel blends at a load of 200 Nm and speed of 1600 rpm.</p>
Full article ">Figure 26
<p>Contours of NOx mass fraction of different fuel blends at a load of 200 Nm and speed of 1800 rpm.</p>
Full article ">Figure 27
<p>Numerical results of NOx emissions with engine speed for all additives.</p>
Full article ">Figure 28
<p>Comparison of numerical and experimental results of BTE with the load for D80B20 and D80B20N150.</p>
Full article ">Figure 29
<p>The compared numerical and experimental results of peak pressure with the engine load for D80B20 and D80B20N150 blends.</p>
Full article ">
12 pages, 6085 KiB  
Article
Demonstration of Polyethylene Nitrous Oxide Catalytic Decomposition Hybrid Thruster with Dual-Catalyst Bed Preheated by Hydrogen Peroxide
by Seungho Lee, Vincent Mario Pierre Ugolini, Eunsang Jung and Sejin Kwon
Aerospace 2025, 12(2), 158; https://doi.org/10.3390/aerospace12020158 - 18 Feb 2025
Abstract
Although various studies on nitrous oxide as a prospective green propellant have been recently explored, a polyethylene nitrous oxide catalytic decomposition hybrid thruster was barely demonstrated due to an inordinately high catalyst preheating time of a heater, which led to the destruction of [...] Read more.
Although various studies on nitrous oxide as a prospective green propellant have been recently explored, a polyethylene nitrous oxide catalytic decomposition hybrid thruster was barely demonstrated due to an inordinately high catalyst preheating time of a heater, which led to the destruction of components. Therefore, hydrogen peroxide was used as a preheatant, a substance to preheat, with a dual-catalyst bed. The thruster with polyethylene (PE) as a fuel, N2O as an oxidizer, H2O2 as the preheatant, Ru/Al2O3 as a catalyst for the oxidizer, and Pt/Al2O3 as a catalyst for the preheatant was arranged. A preheatant supply time of 10 s with a maximum catalyst bed temperature of more than 500 °C and without combustion and an oxidizer supply time of 20 s with a burning time of approximately 15 s were decided. Because the catalyst bed upstream part for decomposing the preheatant was far from the post-combustion chamber, the post-combustion chamber pressure increased and the preheatant mass flow rate decreased after a hard start during the preheatant supply time. Moreover, because the catalyst bed upstream part primarily contributed to preheating, the maximum catalyst bed temperature was less than the decomposition temperature of the preheatant during the preheatant supply time. Additionally, because the catalyst bed downstream part for decomposing the oxidizer was far from the post-combustion chamber, the post-combustion chamber pressure decreased and then increased during a transient state in the oxidizer supply time. Full article
(This article belongs to the Special Issue Green Propellants for In-Space Propulsion)
Show Figures

Figure 1

Figure 1
<p>Blockages of catalyst bed void, distributor holes, and fuel port due to PE melted by heater.</p>
Full article ">Figure 2
<p>Schematic of PE N<sub>2</sub>O catalytic decomposition hybrid thruster with dual-catalyst bed preheated by H<sub>2</sub>O<sub>2</sub>.</p>
Full article ">Figure 3
<p>Schematic of injector manifold.</p>
Full article ">Figure 4
<p>Catalysts for preheatant and oxidizer.</p>
Full article ">Figure 5
<p>Schematic of experimental setup.</p>
Full article ">Figure 6
<p>Schematic of sequences of combustion tests.</p>
Full article ">Figure 7
<p>Combustion test O2-1.</p>
Full article ">Figure 8
<p>Fuel before and after combustion test O2-1.</p>
Full article ">Figure 9
<p>Pressure and mass flow rate according to time of combustion test O2-1.</p>
Full article ">Figure 10
<p>Temperatures according to time of combustion test O2-1.</p>
Full article ">
Back to TopTop