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

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline

Search Results (202)

Search Parameters:
Keywords = boron carbide

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
13 pages, 6097 KiB  
Article
Reduced Graphene Oxide Reinforces Boron Carbide with High-Pressure and High-Temperature Sintering
by Xiaonan Wang, Dianzhen Wang, Kaixuan Rong, Qiang Tao and Pinwen Zhu
Materials 2024, 17(23), 5838; https://doi.org/10.3390/ma17235838 - 28 Nov 2024
Viewed by 46
Abstract
Introducing a second phase has been an effective way to solve the brittleness of boron carbide (B4C) for its application. Though reduced graphene oxide (rGO) is an ideal candidate for reinforcing the B4C duo’s two-dimensional structure and excellent mechanical [...] Read more.
Introducing a second phase has been an effective way to solve the brittleness of boron carbide (B4C) for its application. Though reduced graphene oxide (rGO) is an ideal candidate for reinforcing the B4C duo’s two-dimensional structure and excellent mechanical properties, the toughness is less than 6 MPa·m1/2, or the hardness is lower than 30 GPa in B4C–graphene composites. A barrier to enhancing toughness is the weak interface strength between rGO and B4C, which limits the bridging and pull-out toughening effects of rGO. In this work, internal stress was introduced using a high-pressure and high-temperature (HPHT) method with B4C–rGO composites. The optimal hardness and toughness values for the B4C-2 vol% rGO composite reached 30.1 GPa and 8.6 MPa·m1/2, respectively. The improvement in toughness was 4 times higher than that of pure B4C. The internal stress in the composite increased gradually from 2.3 GPa to 3.3 GPa with an increase in rGO content from 1 vol% to 3 vol%. Crack deflection, bridging, and rGO pull-out are responsible for the improvement in toughness. Moreover, the high internal stress contributed to the formation of good interface strength by embedding rGO into the B4C matrix particles, which further enhanced the dissipation of the crack energy during the pull-out process and led to high toughness. This work provides new insights into synthesizing high-toughness B4C matrix composites. Full article
Show Figures

Figure 1

Figure 1
<p>Relative density of the B<sub>4</sub>C–rGO composites with different rGO contents synthesized at different temperatures.</p>
Full article ">Figure 2
<p>XRD patterns of as-sintered B<sub>4</sub>C–rGO composites. (<b>a</b>) Samples with different rGO contents synthesized at 5 GPa/1400 °C/10 min; (<b>b</b>) B<sub>4</sub>C-2 <span class="html-italic">vol%</span> rGO composites synthesized at 1400–1600 °C.</p>
Full article ">Figure 3
<p>Hardness and fracture toughness profiles of B<sub>4</sub>C–rGO composites. (<b>a</b>–<b>c</b>) Influence of loading force and temperature on the Vickers hardness for samples with different rGO contents; the insets in (<b>a</b>–<b>c</b>) are optical microscopic images of the Vickers indentation at a 9.8 N load; (<b>d</b>) fracture toughness versus rGO content and temperature.</p>
Full article ">Figure 4
<p>SEM images of fracture surface for samples with different rGO contents ((<b>a</b>) 1 <span class="html-italic">vol%</span>, (<b>c</b>) 2 <span class="html-italic">vol%</span>, and (<b>e</b>) 3 <span class="html-italic">vol%</span>) synthesized at 5 GPa/1500 °C/10 min; (<b>b</b>,<b>d</b>,<b>f</b>) shown at a higher magnification.</p>
Full article ">Figure 5
<p>(<b>a</b>) SEM image of Vickers hardness indentation produced at a load of 9.8 N on the polished surface of B<sub>4</sub>C-2 <span class="html-italic">vol%</span> rGO sample synthesized at 5 GPa/1500 °C/10 min; (<b>b</b>) pull-out of rGO; (<b>c</b>) crack bridging and deflection; (<b>d</b>) EDS hierarchical image of element distribution in (<b>c</b>); (<b>e</b>) Raman spectra of B<sub>4</sub>C-2 <span class="html-italic">vol%</span> rGO sample, taken from a crack caused by Vickers indentation; (<b>f</b>) Raman spectra of un-sintered mixed powers.</p>
Full article ">Figure 6
<p>(<b>a</b>,<b>d</b>) TEM images of B<sub>4</sub>C-<span class="html-italic">2 vol%</span> rGO composite synthesized at 5 GPa/1500 °C/10 min; (<b>b</b>,<b>e</b>) HRTEM image of the square area in (<b>a</b>,<b>d</b>), respectively; (<b>c</b>,<b>f</b>) higher magnification of (<b>b</b>,<b>e</b>).</p>
Full article ">Figure 7
<p>C 1s XPS spectra of GO powders (<b>a</b>) and rGO after HPHT sintering at 5 GPa/1500 °C/10 min (<b>b</b>).</p>
Full article ">Figure 8
<p>Raman spectra of un-sintered mixed powers and B<sub>4</sub>C–rGO composites with different rGO contents synthesized at 5 GPa/1500 °C/10 min.</p>
Full article ">Figure 9
<p>Plot of the Vickers hardness and fracture toughness of this work in comparison with previous reports.</p>
Full article ">
22 pages, 6414 KiB  
Article
Experimental Investigation and Machine Learning Modeling of Tribological Characteristics of AZ31/B4C/GNPs Hybrid Composites
by Dhanunjay Kumar Ammisetti, Bharat Kumar Chigilipalli, Baburao Gaddala, Ravi Kumar Kottala, Radhamanohar Aepuru, T. Srinivasa Rao, Seepana Praveenkumar and Ravinder Kumar
Crystals 2024, 14(12), 1007; https://doi.org/10.3390/cryst14121007 - 21 Nov 2024
Viewed by 401
Abstract
In this study, the AZ31 hybrid composites reinforced with boron carbide (B4C) and graphene nano-platelets (GNPs) are prepared by the stir casting method. The main aim of the study is to study the effect of various wear parameters (reinforcement percentage (R), [...] Read more.
In this study, the AZ31 hybrid composites reinforced with boron carbide (B4C) and graphene nano-platelets (GNPs) are prepared by the stir casting method. The main aim of the study is to study the effect of various wear parameters (reinforcement percentage (R), applied load (L), sliding distance (D), and velocity (V)) on the wear characteristics (wear rate (WR)) of the AZ91/B4C/GNP composites. Experiments are designed using the Taguchi technique, and it was determined that load (L) is the most significant parameter affecting WR, followed by D, R, and V. The wear mechanisms under conditions of maximum and minimum wear rates are examined using SEM analysis of the worn-out surfaces of the specimens. From the result analysis on the WR, the ideal conditions for achieving the lowest WR are R = 4 wt.%, L = 15 N, V = 3 m/s, and D = 500 m. Machine learning (ML) models, including linear regression (LR), polynomial regression (PR), random forest (RF), and Gaussian process regression (GPR), are implemented to develop a reliable prediction model that forecasts output responses in accordance with input variables. A total of 90% of the experimental data points were used to train and 10% to evaluate the models. The PR model exceeded the accuracy of other models in predicting WR, with R2 = 0.953, MSE = 0.011, RMSE = 0.103, and COF with R2 = 0.937, MSE = 0.013, and RMSE = 0.114, respectively. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart.</p>
Full article ">Figure 2
<p>SEM image of (<b>a</b>) GNPs and (<b>b</b>) B<sub>4</sub>C; EDS image of (<b>c</b>) GNPs and (<b>d</b>) B<sub>4</sub>C; XRD image of (<b>e</b>) GNPs and (<b>f</b>) B<sub>4</sub>C.</p>
Full article ">Figure 2 Cont.
<p>SEM image of (<b>a</b>) GNPs and (<b>b</b>) B<sub>4</sub>C; EDS image of (<b>c</b>) GNPs and (<b>d</b>) B<sub>4</sub>C; XRD image of (<b>e</b>) GNPs and (<b>f</b>) B<sub>4</sub>C.</p>
Full article ">Figure 3
<p>(<b>a</b>) Wear testing machine. (<b>b</b>) Experimental setup.</p>
Full article ">Figure 4
<p>SEM microstructures of (<b>a</b>) AZ31 + 1 wt.% graphene + 1 wt.% B<sub>4</sub>C; (<b>b</b>) AZ31 + 1 wt.% graphene + 2 wt.% B<sub>4</sub>C; and (<b>c</b>) AZ31 + 1 wt.% graphene + 3 wt.% B<sub>4</sub>C.</p>
Full article ">Figure 5
<p>Effect of various factors on WR (means data).</p>
Full article ">Figure 6
<p>Effect of various factors on WR (S/N ratios data).</p>
Full article ">Figure 7
<p>Interaction plot for means.</p>
Full article ">Figure 8
<p>Residual plots for WR.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) High worn surfaces. (<b>c</b>,<b>d</b>) Low worn out surfaces.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>,<b>b</b>) High worn surfaces. (<b>c</b>,<b>d</b>) Low worn out surfaces.</p>
Full article ">Figure 10
<p>Regression plots for WR data with (<b>a</b>) LR, (<b>b</b>) PR, (<b>c</b>) RF, and (<b>d</b>) GPR. (<b>e</b>) Comparison plot for training and testing of LR, PR, RF, and GPR techniques.</p>
Full article ">Figure 11
<p>Regression plots for COF data with (<b>a</b>) LR, (<b>b</b>) PR, (<b>c</b>) RF, and (<b>d</b>) GPR. (<b>e</b>) Comparison plot for training and testing of LR, PR, RF, and GPR techniques.</p>
Full article ">
19 pages, 7565 KiB  
Article
Improving Mechanical Properties of Low-Quality Pure Aluminum by Minor Reinforcement with Fine B4C Particles and T6 Heat Treatment
by Maxat Abishkenov, Ilgar Tavshanov, Nikita Lutchenko, Nursultan Amanzholov, Daniyar Kalmyrzayev, Zhassulan Ashkeyev, Kayrosh Nogaev, Saltanat Kydyrbayeva and Assylbek Abdirashit
Appl. Sci. 2024, 14(23), 10773; https://doi.org/10.3390/app142310773 - 21 Nov 2024
Viewed by 422
Abstract
Pure aluminum, due to its inherent low strength and softness, is unsuitable for most structural applications. However, unlike many aluminum alloys, pure aluminum exhibits high ductility and is often free from expensive alloying elements. This makes it a promising candidate for minor reinforcement [...] Read more.
Pure aluminum, due to its inherent low strength and softness, is unsuitable for most structural applications. However, unlike many aluminum alloys, pure aluminum exhibits high ductility and is often free from expensive alloying elements. This makes it a promising candidate for minor reinforcement to produce cost-effective composites with an optimal balance of strength and ductility. This study assesses the possibility of improving the mechanical performance of pure aluminum specimens by minor reinforcement (~0.36 wt. %) with fine B4C particles and T6 heat treatment. The composites were obtained using ultrasonic-assisted stir casting and were characterized by assessing their density, microhardness, yield strength (YS), ultimate tensile strength (UTS), and elongation. Light microscopy (LM), scanning electron microscopy (SEM), energy-dispersive spectroscopy (EDS), and X-ray diffraction (XRD) tests were conducted to investigate the presence and distribution of reinforcing particles in the Al matrix. Minor reinforcement of ~0.5–2 μm with B4C particles without/with subsequent T6 heat treatment resulted in an increase in microhardness by 71.45% and 143.37% and UTS by 71.05% and 140.16%, respectively, while the elongation values of the specimens decreased to 51.98% and 42.38%, respectively, compared with the adopted initial matrix Al specimen. Full article
(This article belongs to the Section Materials Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>Illustration of the stages of the AMCs fabrication process.</p>
Full article ">Figure 2
<p>Illustration of the specimen mass measurement process in the MH-300A densitometer for experimental density determination.</p>
Full article ">Figure 3
<p>Illustration of Vickers microhardness measurement using the HVT-1000A microhardness tester (Laizhou Laihua Testing Instrument Factory, Laizhou, China).</p>
Full article ">Figure 4
<p>Dimensions (in mm) according to ASTM E8 [<a href="#B31-applsci-14-10773" class="html-bibr">31</a>] and appearance of prepared and tested tensile test specimens.</p>
Full article ">Figure 5
<p>SEM images and EDS spectra of boron carbide (B<sub>4</sub>C) and potassium hexafluorotitanate (K<sub>2</sub>TiF<sub>6</sub>) particles.</p>
Full article ">Figure 6
<p>LM images of the S1, S1-HT, S2, and S2-HT specimens.</p>
Full article ">Figure 7
<p>SEM images and EDS spectra of the S1, S1-HT, S2, and S2-HT specimens.</p>
Full article ">Figure 8
<p>Combined XRD patterns of the S1, S1-HT, S2, and S2-HT specimens.</p>
Full article ">Figure 9
<p>Combined XRD patterns of S2 and S2-HT specimens (fragmented for B<sub>4</sub>C).</p>
Full article ">Figure 10
<p>Density and porosity of S1, S1-HT, S2 and S2-HT specimens.</p>
Full article ">Figure 11
<p>Measured microhardness values (HV0.1) of the S1, S1-HT, S2, and S2-HT specimens.</p>
Full article ">Figure 12
<p>Stress–strain curves (<b>a</b>) and the histogram comparing the values of yield strength (YS), ultimate tensile strength (UTS), and elongation (<b>b</b>) of the S1, S1-HT, S2, and S2-HT specimens.</p>
Full article ">
17 pages, 39089 KiB  
Article
Electronic and Optical Properties of 2D Heterostructure Bilayers of Graphene, Borophene and 2D Boron Carbides from First Principles
by Lu Niu, Oliver J. Conquest, Carla Verdi and Catherine Stampfl
Nanomaterials 2024, 14(20), 1659; https://doi.org/10.3390/nano14201659 - 16 Oct 2024
Viewed by 815
Abstract
In the present work the atomic, electronic and optical properties of two-dimensional graphene, borophene, and boron carbide heterojunction bilayer systems (Graphene–BC3, Graphene–Borophene and Graphene–B4C3) as well as their constituent monolayers are investigated on the basis of first-principles [...] Read more.
In the present work the atomic, electronic and optical properties of two-dimensional graphene, borophene, and boron carbide heterojunction bilayer systems (Graphene–BC3, Graphene–Borophene and Graphene–B4C3) as well as their constituent monolayers are investigated on the basis of first-principles calculations using the HSE06 hybrid functional. Our calculations show that while borophene is metallic, both monolayer BC3 and B4C3 are indirect semiconductors, with band-gaps of 1.822 eV and 2.381 eV as obtained using HSE06. The Graphene–BC3 and Graphene–B4C3 bilayer heterojunction systems maintain the Dirac point-like character of graphene at the K-point with the opening of a very small gap (20–50 meV) and are essentially semi-metals, while Graphene–Borophene is metallic. All bilayer heterostructure systems possess absorbance in the visible region where the resonance frequency and resonance absorption peak intensity vary between structures. Remarkably, all heterojunctions support plasmons within the range 16.5–18.5 eV, while Graphene–B4C3 and Graphene–Borophene exhibit a π-type plasmon within the region 4–6 eV, with the latter possessing an additional plasmon at the lower energy of 1.5–3 eV. The dielectric tensor for Graphene–B4C3 exhibits complex off-diagonal elements due to the lower P3 space group symmetry indicating it has anisotropic dielectric properties and could exhibit optically active (chiral) effects. Our study shows that the two-dimensional heterostructures have desirable optical properties broadening the potential applications of the constituent monolayers. Full article
Show Figures

Figure 1

Figure 1
<p>Optimized atomic structures of BC<sub>3</sub>, borophene, B<sub>4</sub>C<sub>3</sub> and graphene. Boron and carbon atoms are denoted by the green and brown spheres, respectively. Borophene has three unique bonds indicated by <math display="inline"><semantics> <msub> <mi>l</mi> <mn>1</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>2</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>l</mi> <mn>3</mn> </msub> </semantics></math>, while B<sub>4</sub>C<sub>3</sub> has four unique bonds indicated by <math display="inline"><semantics> <msub> <mi>l</mi> <mn>4</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>5</mn> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>l</mi> <mn>6</mn> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mi>l</mi> <mn>7</mn> </msub> </semantics></math>. The unit cells are highlighted in orange.</p>
Full article ">Figure 2
<p>Band structure of the four monolayer systems, BC<sub>3</sub>, B<sub>4</sub>C<sub>3</sub>, borophene and graphene as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 3
<p>Top view of the optimized atomic structures of Graphene–BC<sub>3</sub>, Graphene–Borophene, and Graphene–B<sub>4</sub>C<sub>3</sub>. Boron and carbon atoms are denoted by the green and brown spheres, respectively. The unit cell is indicated by the orange parallelogram.</p>
Full article ">Figure 4
<p>Charge density difference (<math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>ρ</mi> <mo>(</mo> <mi mathvariant="bold">r</mi> <mo>)</mo> </mrow> </semantics></math>, calculated using Equation (<a href="#FD10-nanomaterials-14-01659" class="html-disp-formula">10</a>)) between the monolayers and the heterostructures. Regions of charge accumulation are shown in yellow and regions of charge depletion are shown in blue. The isosurface level is <math display="inline"><semantics> <mrow> <mn>1.5</mn> <mo>×</mo> <msup> <mn>10</mn> <mrow> <mo>−</mo> <mn>4</mn> </mrow> </msup> </mrow> </semantics></math> a<sub>0</sub><sup>−3</sup> and the top layer is always graphene. The unit cell is indicated by the orange parallelogram.</p>
Full article ">Figure 5
<p>Band structure and total DOS for Graphene–BC<sub>3</sub> as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 6
<p>Band structure and total DOS for Graphene–Borophene as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 7
<p>Band structure and total DOS for Graphene–B<sub>4</sub>C<sub>3</sub> as calculated using the PBE (blue) and HSE06 (orange) functionals. The Fermi level is indicated by the purple dashed line.</p>
Full article ">Figure 8
<p><b>Left</b>: schematic of the Schottky-Mott model showing the valence and conduction band energies, the Fermi energy, and <span class="html-italic">n</span>-type and <span class="html-italic">p</span>-type Schottky barriers labelled <math display="inline"><semantics> <msub> <mi>E</mi> <mi>V</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>E</mi> <mi>C</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mi>E</mi> <mi>F</mi> </msub> </semantics></math>, <math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>n</mi> </msub> </semantics></math>, and <math display="inline"><semantics> <msub> <mo>Φ</mo> <mi>p</mi> </msub> </semantics></math>, respectively. <b>Right</b>: Example for the HSE06 calculated Graphene–B<sub>4</sub>C<sub>3</sub> heterojunction showing the determined Schottky barrier height from the band structure.</p>
Full article ">Figure 9
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–Borophene heterostructure.</p>
Full article ">Figure 10
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–BC<sub>3</sub> heterostructure.</p>
Full article ">Figure 11
<p>The real and imaginary parts of the in-plane and out-of-plane dielectric function as a function of photon energy as calculated using the PBE and HSE06 functionals for the Graphene–B<sub>4</sub>C<sub>3</sub> heterostructure.</p>
Full article ">Figure 12
<p>The in-plane and out-of-plane adsorption coefficient (upper) and energy loss spectrum (lower) as a function of photon energy as calculated using the HSE06 functional.</p>
Full article ">Figure 13
<p>The in-plane and out-of-plane reflectivity (upper) and refractive index (lower) as a function of photon energy as calculated using the HSE06 functional.</p>
Full article ">
11 pages, 4809 KiB  
Article
Binderless Polycrystalline Cubic Boron Nitride Sintered Compacts for Machining of Cemented Carbides
by Alexander S. Osipov, Piotr Klimczyk, Igor A. Petrusha, Yurii O. Melniichuk, Lucyna Jaworska, Kinga Momot and Yuliia Rumiantseva
Ceramics 2024, 7(4), 1477-1487; https://doi.org/10.3390/ceramics7040095 - 13 Oct 2024
Viewed by 687
Abstract
High-purity, superhard, binderless polycrystalline cubic boron nitride (BL-PCBN) was obtained by direct hBN to cBN transformation in a toroid-type high-pressure apparatus at a pressure of 8.0 GPa and temperature of 2250 °C (HPHT-DCS; high-pressure, high-temperature direct conversion sintering). X-ray diffraction analysis revealed a [...] Read more.
High-purity, superhard, binderless polycrystalline cubic boron nitride (BL-PCBN) was obtained by direct hBN to cBN transformation in a toroid-type high-pressure apparatus at a pressure of 8.0 GPa and temperature of 2250 °C (HPHT-DCS; high-pressure, high-temperature direct conversion sintering). X-ray diffraction analysis revealed a prominent [111] axial texture in the sintered material when the axis was oriented perpendicular to the end surface of the sample. Vickers hardness tests conducted at a load of 49 N showed that BL-PCBN possessed an exceptional hardness value of 63.4 GPa. Finally, cutting tools made of BL-PCBN and SN-PCBN (Si3N4-doped cBN-based composite) reference materials were tested during the turning of a cemented tungsten carbide workpiece. The results of the cutting tests demonstrated that the wear resistance of the BL-PCBN material obtained with the HPHT-DCS process is 1.5–1.9 times higher compared to the conventional SN-PCBN material, suggesting its significant potential for industrial application. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>HPC view in the axial section: (<b>a</b>) central part of the assembly (half section) before compression; (<b>b</b>) HPC in compressed state (half section); (<b>c</b>) HPA-T30 appearance; (1) pyrophyllite heat-insulating ring, (2) axial heater (ZrO<sub>2</sub> + graphite), (3) pressed graphite disc, (4) CsCl + graphite, (5) graphite heater, (6) specimen, (7) outer supporting punch, (8) WC-Co anvil of HPA, (9) fastening steel ring, (10) CaCO<sub>3</sub> internal gasket; (11) external gasket (pressed CaCO<sub>3</sub> + binder).</p>
Full article ">Figure 2
<p>Justification of the choice of HPHT-DCS parameters for the present study (frame) against the background of the results of previous studies on complete hBN→cBN conversion (<b>a</b>) with the formation of an ultra-fine-grain monophase structure of the polycrystalline cBN (<b>b</b>).</p>
Full article ">Figure 3
<p>Turning of cemented carbide bushing of the WC-Co system (Co 15 wt.%) using the BL-PCBN (RNMN 09T300F) cutting insert.</p>
Full article ">Figure 4
<p>Initial hBN powder: (<b>a</b>) SEM image of the powder morphology; (<b>b</b>) XRD pattern (the Miller indices of the planes in the hBN lattice are given only for the most intense reflections).</p>
Full article ">Figure 5
<p>Translucent BL-PCBN obtained from hBN powder by the HPHT-DCS method: (<b>a</b>) sample view (sample thickness 3.6 mm); (<b>b</b>) XRD pattern of the sample.</p>
Full article ">Figure 6
<p>Change of flank wear width with cutting speed during the turning of the WC-15Co alloy (depth of cut—0.2 mm; feed—0.1 mm/rev; cutting length—75 m).</p>
Full article ">Figure 7
<p>Cutting edges of SN-PCBN (<b>a</b>,<b>c</b>) and BL-PCBN (<b>b</b>,<b>d</b>) tools after the turning of the WC-15Co material at 37 m/min (<b>a</b>,<b>b</b>) and 60 m/min (<b>c</b>,<b>d</b>). The distance between the dashed lines corresponds to the flank wear.</p>
Full article ">Figure 8
<p>The turning performance of WC-15Co alloy cutting: depth of cut—0.2 mm; feed—0.1 mm/rev; cutting length—75 m; flank wear width—0.4 mm.</p>
Full article ">
14 pages, 11777 KiB  
Article
Increasing the Wear and Corrosion Resistance of a CP-Ti Surface by Plasma Electrolytic Borocarburizing and Polishing
by Marina A. Volosova, Sergei A. Kusmanov, Ivan V. Tambovskiy, Tatiana L. Mukhacheva, Artem P. Mitrofanov, Igor V. Suminov and Sergey N. Grigoriev
Surfaces 2024, 7(4), 824-837; https://doi.org/10.3390/surfaces7040054 - 7 Oct 2024
Viewed by 963
Abstract
The paper examines the possibility of increasing the wear and corrosion resistance of a CP-Ti surface by duplex plasma electrolytic treatment (borocarburizing and polishing). The structure and composition of diffusion layers, their microhardness, surface morphology and roughness, wear resistance during dry friction and [...] Read more.
The paper examines the possibility of increasing the wear and corrosion resistance of a CP-Ti surface by duplex plasma electrolytic treatment (borocarburizing and polishing). The structure and composition of diffusion layers, their microhardness, surface morphology and roughness, wear resistance during dry friction and corrosion resistance in Ringer’s solution were studied. The formation of a surface-hardened layer up to 200 μm thick with a microhardness of up to 950 HV, including carbides and a solid solution of boron and carbon, is shown. Subsequent polishing makes it possible to reduce surface roughness and remove weak areas of the porous oxide layer, which are formed during high-temperature oxidation in aqueous electrolyte vapor during borocarburizing. Changing the morphology and structural-phase composition of the CP-Ti surface helps reduce weight wear by a factor of three (the mode of frictional interaction changes from microcutting to oxidative wear) and corrosion current density by a factor of four after borocarburizing in a solution of boric acid, glycerin and ammonium chloride at 950 °C for 5 min and subsequent polishing in an ammonium fluoride solution at a voltage of 250 V for 3 min. Full article
Show Figures

Figure 1

Figure 1
<p>PEBC and PEP installation scheme: 1—ventilation duct; 2—protective screen of the working chamber; 3—linear drive; 4—workpiece–electrode (anode); 5—cylindrical cell–electrode (cathode); 6—working chamber; 7—valve with electric drive; 8—flow meter; 9—heat exchanger; 10—pump; 11—water filter.</p>
Full article ">Figure 2
<p>Friction scheme and unit: 1—cylindrical sample; 2—counter body; 3, 4—shaft; 5—crank; 6—pneumatic cylinder; 7—guides; 8—table; 9—strain gauges.</p>
Full article ">Figure 3
<p>Surface morphology of CP-Ti samples after PEBC at different temperatures: (<b>a</b>) 800 °C; (<b>b</b>) 850 °C; (<b>c</b>) 900 °C; (<b>d</b>) 950 °C.</p>
Full article ">Figure 4
<p>X-ray diffraction pattern of the surface of CP-Ti samples after PEBC at different temperatures.</p>
Full article ">Figure 5
<p>SEM image of the cross section of CP-Ti samples after PEBC at different temperatures: (<b>a</b>) 800 °C; (<b>b</b>) 850 °C; (<b>c</b>) 900 °C; (<b>d</b>) 950 °C. 1—oxide layer; 2—PEBC layer; 3—initial structure.</p>
Full article ">Figure 6
<p>Microhardness of the surface layer of CP-Ti samples after PEBC at different temperatures.</p>
Full article ">Figure 7
<p>Morphology of friction tracks of CP-Ti samples before (<b>a</b>) and after PEBC at different temperatures: (<b>b</b>) 800 °C; (<b>c</b>) 850 °C; (<b>d</b>) 900 °C; (<b>e</b>) 950 °C.</p>
Full article ">Figure 8
<p>Surface morphology of CP-Ti samples before (<b>a</b>) and after PEP at different times: (<b>b</b>) 1 min; (<b>c</b>) 3 min; (<b>d</b>) 5 min.</p>
Full article ">Figure 9
<p>The microhardness of the surface layer of CP-Ti samples before (0 min) and after PEP at different times.</p>
Full article ">Figure 10
<p>The X-ray diffraction pattern of the surface of CP-Ti samples before (0 min) and after PEP at different times.</p>
Full article ">Figure 11
<p>Morphology of friction tracks of CP-Ti samples before (<b>a</b>) and after PEP at different times: (<b>b</b>) 1 min; (<b>c</b>) 3 min; (<b>d</b>) 5 min.</p>
Full article ">
12 pages, 11316 KiB  
Article
Toughening Mechanism in Nanotwinned Boron Carbide: A Molecular Dynamics Study
by Hongchi Zhang, Yesheng Zhong, Xiaoliang Ma, Lin Yang, Xiaodong He and Liping Shi
Nanomaterials 2024, 14(18), 1493; https://doi.org/10.3390/nano14181493 - 14 Sep 2024
Viewed by 716
Abstract
Boron carbide ceramics are potentially ideal candidates for lightweight bulletproof armor, but their use is currently limited by their low fracture toughness. Recent experimental results have shown that sintered samples with high twin densities exhibit high fracture toughness, but the toughening mechanism and [...] Read more.
Boron carbide ceramics are potentially ideal candidates for lightweight bulletproof armor, but their use is currently limited by their low fracture toughness. Recent experimental results have shown that sintered samples with high twin densities exhibit high fracture toughness, but the toughening mechanism and associated crack propagation process of nanotwinned boron carbide at the atomic scale remain a mystery. Reported here are molecular dynamics simulations with a reactive force field potential to investigate how nanoscale twins affect the fracture toughness of boron carbide ceramics. The results show that the strength disparity on either side of a twin boundary is the fundamental reason for the toughening effect; the twin boundary impedes crack propagation only when the crack moves to a region of higher fracture strength. The fracture toughness of nanotwinned boron carbide is greatly affected by the angle between the twin boundary and the prefabricated crack. At an angle of 120°, the twin boundary provides the maximum toughening effect, enhancing the toughness by 32.72%. Moreover, phase boundaries—another common structure in boron carbide ceramics—have no toughening effect. This study provides new insights into the design of boron carbide ceramics with high fracture toughness. Full article
(This article belongs to the Special Issue Theoretical Calculation Study of Nanomaterials: 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Left: atomic structures of (<b>a</b>) twin boundary (TB) and (<b>b</b>) phase boundary (PB). Right: simulation models for fracture toughness of (<b>c</b>) nanotwinned (NT) B<sub>4</sub>C and (<b>d</b>) PB B<sub>4</sub>C.</p>
Full article ">Figure 2
<p>Stress–strain relationships of single-crystal (SC) and NT B<sub>4</sub>C at several typical included angles.</p>
Full article ">Figure 3
<p>Changes in von Mises stress distribution with crack propagation for NT (first row) and SC (second row) B<sub>4</sub>C models at angles of (<b>a</b>) 90°, (<b>b</b>) 40°, and (<b>c</b>) 140°.</p>
Full article ">Figure 4
<p>Variation in fracture energy of NT and SC B<sub>4</sub>C with included angle.</p>
Full article ">Figure 5
<p>Stress–strain relationship of SC-120 model and NT-120 model.</p>
Full article ">Figure 6
<p>Stress–strain relationships of SC and PB B<sub>4</sub>C at several typical included angles.</p>
Full article ">Figure 7
<p>Changes in von Mises stress distribution with crack propagation for PB (first row) and SC (second row) B<sub>4</sub>C models at angles of (<b>a</b>) 90°, (<b>b</b>) 40°, and (<b>c</b>) 140°.</p>
Full article ">Figure 8
<p>Variation in fracture energy of PB and SC B<sub>4</sub>C with included angle.</p>
Full article ">
20 pages, 7965 KiB  
Article
Optimization of Dry Sliding Wear in Hot-Pressed Al/B4C Metal Matrix Composites Using Taguchi Method and ANN
by Sandra Gajević, Slavica Miladinović, Onur Güler, Serdar Özkaya and Blaža Stojanović
Materials 2024, 17(16), 4056; https://doi.org/10.3390/ma17164056 - 15 Aug 2024
Cited by 3 | Viewed by 967
Abstract
The presented study investigates the effects of weight percentages of boron carbide reinforcement on the wear properties of aluminum alloy composites. Composites were fabricated via ball milling and the hot extrusion process. During the fabrication of composites, B4C content was varied [...] Read more.
The presented study investigates the effects of weight percentages of boron carbide reinforcement on the wear properties of aluminum alloy composites. Composites were fabricated via ball milling and the hot extrusion process. During the fabrication of composites, B4C content was varied (0, 5, and 10 wt.%), as well as milling time (0, 10, and 20 h). Microstructural observations with SEM microscopy showed that with an increase in milling time, the distribution of B4C particles is more homogeneous without agglomerates, and that an increase in wt.% of B4C results in a more uniform distribution with distinct grain boundaries. Taguchi and ANOVA analyses are applied in order to investigate how parameters like particle content of B4C, normal load, and milling time affect the wear properties of AA2024-based composites. The ANOVA results showed that the most influential parameters on wear loss and coefficient of friction were the content of B4C with 51.35% and the normal load with 45.54%, respectively. An artificial neural network was applied for the prediction of wear loss and the coefficient of friction. Two separate networks were developed, both having an architecture of 3-10-1 and a tansig activation function. By comparing the predicted values with the experimental data, it was demonstrated that the well-trained feed-forward-back propagation ANN model is a powerful tool for predicting the wear behavior of Al2024-B4C composites. The developed models can be used for predicting the properties of Al2024-B4C composite powders produced with different reinforcement ratios and milling times. Full article
Show Figures

Figure 1

Figure 1
<p>The production steps followed in the manufacturing of AA2024-B<sub>4</sub>C composites.</p>
Full article ">Figure 2
<p>Schematics of basic ANN.</p>
Full article ">Figure 3
<p>Main effects plot for S/N ratio for the (<b>a</b>) wear loss and (<b>b</b>) coefficient of friction.</p>
Full article ">Figure 4
<p>Regression coefficients for (<b>a</b>) wear loss and (<b>b</b>) CoF.</p>
Full article ">Figure 5
<p>Comparative presentation of the results of the experiment, regression model, and ANN model for (<b>a</b>) wear loss and (<b>b</b>) CoF.</p>
Full article ">Figure 6
<p>SEM images of the microstructure of (<b>a</b>) AA2024 alloy, (<b>b</b>) AA2024 + 5% B<sub>4</sub>C, and (<b>c</b>) AA2024 + 10% B<sub>4</sub>C composites with the RD values.</p>
Full article ">Figure 7
<p>SEM images of the microstructure of AA2024 + 5% B<sub>4</sub>C obtained with (<b>a</b>) 0 h, (<b>b</b>) 10 h, and (<b>c</b>) 20 h milling with the RD values.</p>
Full article ">Figure 8
<p>Wear tracks SEM images of the samples obtained under different parameters.</p>
Full article ">
25 pages, 31123 KiB  
Article
Empirical Investigation of Properties for Additive Manufactured Aluminum Metal Matrix Composites
by Shuang Bai and Jian Liu
Appl. Mech. 2024, 5(3), 450-474; https://doi.org/10.3390/applmech5030026 - 11 Jul 2024
Viewed by 983
Abstract
Laser additive manufacturing with mixed powders of aluminum alloy and silicon carbide (SiC) or boron carbide (B4C) is investigated in this experiment. With various mixing ratios of SiC/Al to form metal matrix composites (MMC), their mechanical and physical properties are empirically [...] Read more.
Laser additive manufacturing with mixed powders of aluminum alloy and silicon carbide (SiC) or boron carbide (B4C) is investigated in this experiment. With various mixing ratios of SiC/Al to form metal matrix composites (MMC), their mechanical and physical properties are empirically investigated. Parameters such as laser power, scan speed, scan pattern, and hatching space are optimized to obtain the highest density for each mixing ratio of SiC/Al. The mechanical and thermal properties are systematically investigated and compared with and without heat treatment. It shows that 2 wt% of SiC obtained the highest strength and Young’s modulus. Graded composite additive manufacturing (AM) of MMC is also fabricated and characterized. Various types of MMC devices, such as heat sink using graded SiC MMC and grid type three-dimensional (3D) neutron collimators using boron carbide (B4C), were also fabricated to demonstrate their feasibility for applications. Full article
Show Figures

Figure 1

Figure 1
<p>AlSi10Mg powder size measurement.</p>
Full article ">Figure 2
<p>Silicon carbide powder size measurement.</p>
Full article ">Figure 3
<p>Characterization of SiC (silicon carbide) powder with EDX (energy-dispersed X-ray spectroscopy).</p>
Full article ">Figure 4
<p>Water immersion density measurement of AlSi10Mg printed cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 5
<p>Surface roughness, Ra, measurement of AlSi10Mg cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 6
<p>Water immersion density measurement of SiC/Al: 2/98 printed cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 7
<p>Surface roughness, Ra, measurement of SiC/Al: 2/98 printed cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 8
<p>Water immersion density measurement of SiC/Al: 5/95 printed cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 9
<p>Surface roughness, Ra, measurement of SiC/Al: 5/95 printed cubic parts (<b>a</b>) matrix test, (<b>b</b>) verification test.</p>
Full article ">Figure 10
<p>(<b>a</b>) Density comparison of printed cubic parts before and after HIP. (<b>b</b>) Surface roughness comparison.</p>
Full article ">Figure 11
<p>Directions of the cross-section for measurement (American Society for Testing and Materials ASTM 52921).</p>
Full article ">Figure 12
<p>Comparison of microstructure and grain boundary of AlSi10Mg cubic samples along different orientations: (<b>Top</b>) etched for 10 min (as built) (<b>Bottom</b>) etched for 2 h (HIP-treated).</p>
Full article ">Figure 13
<p>Comparison of microstructure and grain boundary of SiC/Al: 2/98 cubic samples along different orientations: (<b>Top</b>) etched for 10 min (as built), (<b>Bottom</b>) etched for 2 h (HIP-treated).</p>
Full article ">Figure 14
<p>Comparison of microstructure and grain boundary of SiC/Al: 5/95 cubic samples along different orientations: (<b>Top</b>) etched for 10 min (as built), (<b>Bottom</b>) etched for 2 h (HIP-treated).</p>
Full article ">Figure 15
<p>Example of standard dog-bone parts printing.</p>
Full article ">Figure 16
<p>Tensile test results of three compositions. (<b>a</b>) Yield strength. (<b>b</b>) Ultimate strength. (<b>c</b>) Young’s modulus. (<b>d</b>) Elongation.</p>
Full article ">Figure 17
<p>(<b>a</b>) Thermal conductivity measurement results and (<b>b</b>) coefficient of thermal expansion measurement results.</p>
Full article ">Figure 18
<p>Phase identification results for AlSi10Mg (<b>a</b>) as-built, (<b>b</b>) HIP-treated.</p>
Full article ">Figure 19
<p>Phase identification for SiC/Al: 2/98 (<b>a</b>) as-built, (<b>b</b>) HIP-treated.</p>
Full article ">Figure 20
<p>Phase identification for SiC/Al: 5/95 (<b>a</b>) as-built, (<b>b</b>) HIP-treated.</p>
Full article ">Figure 20 Cont.
<p>Phase identification for SiC/Al: 5/95 (<b>a</b>) as-built, (<b>b</b>) HIP-treated.</p>
Full article ">Figure 21
<p>Electron backscatter diffraction (EBSD) analysis results of AlSi10Mg (as-built) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map.</p>
Full article ">Figure 22
<p>EBSD analysis results of AlSi10Mg (HIP-treated) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map.</p>
Full article ">Figure 23
<p>EDS mapping images of the cross-section of AlSi10Mg part (as-built).</p>
Full article ">Figure 24
<p>EDS quantitative analysis of AlSi10Mg part (as-built).</p>
Full article ">Figure 25
<p>EBSD analysis results of SiC/Al: 2/98 (as-built) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map. Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 26
<p>EBSD analysis result of SiC/Al: 2/98 (HIP-treated) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map. Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 27
<p>EDS mapping images of the cross-section of SiC/Al: 2/98 part (as-built). Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 27 Cont.
<p>EDS mapping images of the cross-section of SiC/Al: 2/98 part (as-built). Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 28
<p>EDS quantitative analysis of SiC/Al: 2/98 part (as-built).</p>
Full article ">Figure 29
<p>EBSD analysis result of SiC/Al: 5/95 (as-built) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map. Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 30
<p>EBSD analysis results of SiC/Al: 5/95 (HIP-treated) (<b>a</b>) Kikuchi band contrast map, (<b>b</b>) phase map. Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 31
<p>EDS mapping images of the cross-section of SiC/Al: 5/95 part (as-built). Irregular shapes are unmelted SiC powders.</p>
Full article ">Figure 32
<p>EDS quantitative analysis of SiC/Al: 5/95 part (as-built).</p>
Full article ">Figure 33
<p>Composition layers of graded printing test sample.</p>
Full article ">Figure 34
<p>Microscopic images of cross-section for five different locations of the graded composition cubic sample.</p>
Full article ">Figure 35
<p>Graded composition AM of multi-functional heat sink (<b>a</b>) 3D modeling, (<b>b</b>) printed heat sink.</p>
Full article ">Figure 36
<p>Neutron scattering instrumentation at the Spallation Neutron Source of ORNL (<b>middle</b>), boron carbide MMC 3D collimator (<b>left</b>), and 2D collimator (<b>right</b>).</p>
Full article ">Figure 37
<p>B<sub>4</sub>C/Al MMC parts made for ORNL neutron scattering instrumentation. These holder and bracket are used to shield neutron scattering from mounting detection components.</p>
Full article ">
48 pages, 14140 KiB  
Review
Cold Spray Technology and Its Application in the Manufacturing of Metal Matrix Composite Materials with Carbon-Based Reinforcements
by Sheng Dai, Mengchao Cui, Jiahui Li and Meng Zhang
Coatings 2024, 14(7), 822; https://doi.org/10.3390/coatings14070822 - 2 Jul 2024
Viewed by 2548
Abstract
Cold spray technology, as an emerging surface engineering technique, effectively prepares hard coatings by high-speed projection of powder materials onto substrates at relatively low temperatures. The principal advantage of this technology lies in its ability to rapidly deposit coatings without significantly altering the [...] Read more.
Cold spray technology, as an emerging surface engineering technique, effectively prepares hard coatings by high-speed projection of powder materials onto substrates at relatively low temperatures. The principal advantage of this technology lies in its ability to rapidly deposit coatings without significantly altering the properties of the substrate or powder materials. Carbon-based materials, especially carbides and diamond, etc., are renowned for their exceptional hardness and thermal stability, which make them indispensable in industrial applications requiring materials with high wear resistance and durability at elevated temperatures. This review elucidates the fundamental principles of cold spray technology, the key components of the equipment, and the properties and applications of hard coatings. The equipment involved primarily includes spray guns, powder feeders, and gas heaters, while the properties of the coatings, such as mechanical strength, corrosion resistance, and tribological performance, are discussed in detail. Moreover, the application of this technology in preparing metal matrix composite (MMC) materials with carbon-based reinforcements, including tungsten carbide, boron carbide, titanium carbide, and diamond, are particularly emphasized, showcasing its potential to enhance the performance of tools and components. Finally, this article outlines the challenges and prospects faced by cold spray technology, highlighting the importance of material innovation and process optimization. This review provides researchers in the fields of materials science and engineering with a comprehensive perspective on the application of cold spray technology in MMC materials with carbon-based reinforcements to drive significant improvements in coating performance and broaden the scope of its industrial applications. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of cold spray technology principle. Reproduced from [<a href="#B44-coatings-14-00822" class="html-bibr">44</a>].</p>
Full article ">Figure 2
<p>Al-Cu-Mg cold-sprayed coating deposited at 350 °C under 3.75 MPa. (<b>a</b>) Splat structure showing different grain structures inside the splats and at the interface. (<b>b</b>) Equiaxed grains at the center of the splats. Reproduced from [<a href="#B49-coatings-14-00822" class="html-bibr">49</a>].</p>
Full article ">Figure 3
<p>TEM images of the subgrains formed in the Al-Cu-Mg-Mn coatings deposited by cold spray using 3.75 MPa at (<b>a</b>) 350 °C and (<b>b</b>) 500 °C. The subgrain size is clearly reduced at lower temperature. Reproduced from [<a href="#B49-coatings-14-00822" class="html-bibr">49</a>].</p>
Full article ">Figure 4
<p>Optical images showing the evolution of porosity in stainless steel 316L cold-sprayed at 900 °C under (<b>a</b>) 5, (<b>b</b>) 6 and (<b>c</b>) 7 MPa. Reproduced from [<a href="#B55-coatings-14-00822" class="html-bibr">55</a>].</p>
Full article ">Figure 5
<p>The long raster direction was made parallel to the crack growth direction, as indicated in (<b>a</b>). (<b>b</b>) Interface compact tension sample, showing the artificial notch created by the shielding plate. (<b>c</b>) Schematic diagram of the four-point bending test, (<b>d</b>) Notched interfacial four-point bend specimen geometry. Reproduced from [<a href="#B56-coatings-14-00822" class="html-bibr">56</a>].</p>
Full article ">Figure 6
<p>Stress–strain curves of cold spray Inconel 718 coatings deposited by different propelling gas types in AS and HT conditions: (<b>a</b>) overview, (<b>b</b>) magnified view of boxed area in (<b>a</b>). Reproduced from [<a href="#B63-coatings-14-00822" class="html-bibr">63</a>].</p>
Full article ">Figure 7
<p>Electrochemical characterization of Al-Al<sub>2</sub>O<sub>3</sub>/Al coatings. (<b>a</b>) Experimental (symbol) and fitting (solid line) complex plane, and (<b>b</b>) Bode phase plots obtained for coated samples for 600 h of immersion in 3.5 wt% NaCl solution at 25 °C. Reproduced from [<a href="#B68-coatings-14-00822" class="html-bibr">68</a>].</p>
Full article ">Figure 8
<p>Polarization curves of the Al-Mg-Si coating and Al-Mg-Si/Al<sub>2</sub>O<sub>3</sub> composite coatings immersed in 3.5 wt% NaCl solution for 1, 5, 12, and 24 h: (<b>a</b>) Al-Mg-Si; (<b>b</b>) Al-Mg-Si/20 vol.% Al<sub>2</sub>O<sub>3</sub>; (<b>c</b>) Al-Mg-Si/40 vol.% Al<sub>2</sub>O<sub>3</sub>; (<b>d</b>) Al-Mg-Si/60 vol.% Al<sub>2</sub>O<sub>3</sub>. Reproduced from [<a href="#B77-coatings-14-00822" class="html-bibr">77</a>].</p>
Full article ">Figure 9
<p>The friction factor of cold gas spray (CGS) and HVOF coating. Reproduced from [<a href="#B90-coatings-14-00822" class="html-bibr">90</a>].</p>
Full article ">Figure 10
<p>Photos of damaged parts before and after CS repair: S-92 helicopter gearbox sump. Reproduced from [<a href="#B118-coatings-14-00822" class="html-bibr">118</a>].</p>
Full article ">Figure 11
<p>Modification technique of adding a new component to a bearing cap through CSAM. (<b>a</b>) Original component, (<b>b</b>) sprayed component, (<b>c</b>) machined component. Reproduced from [<a href="#B45-coatings-14-00822" class="html-bibr">45</a>].</p>
Full article ">Figure 12
<p>Repairing steps for CSAM. (<b>a</b>) Pre-machining of the damaged zone, (<b>b</b>) material deposition, (<b>c</b>) post-machining on the back-filling material, and (<b>d</b>) performance testing. Reproduced from [<a href="#B45-coatings-14-00822" class="html-bibr">45</a>].</p>
Full article ">Figure 13
<p>Comparing the deteriorated parts and parts restored by CSAM: (<b>a</b>) inner bore surface of a navy valve actuator and (<b>b</b>) UH-60 helicopter gearbox sump. Reproduced from [<a href="#B118-coatings-14-00822" class="html-bibr">118</a>,<a href="#B121-coatings-14-00822" class="html-bibr">121</a>].</p>
Full article ">Figure 14
<p>Average coefficients of friction (CoFs) of Cu-MoS<sub>2</sub>, Cu-MoS<sub>2</sub>-WC, and pure Cu in dry nitrogen. Reproduced from [<a href="#B124-coatings-14-00822" class="html-bibr">124</a>].</p>
Full article ">Figure 15
<p>EBSD Euler angle maps of the samples in (<b>a</b>) as-sprayed condition and after heat treatment at (<b>b</b>) 200 °C, (<b>c</b>) 300 °C, (<b>d</b>) 400 °C, and (<b>e</b>) 500 °C in XZ-plane. (<b>f</b>) Euler angle image of as-received Al powder particle. (N.B., particles with gray contrast in panels (<b>a</b>–<b>e</b>) are B<sub>4</sub>C particles). Reproduced from [<a href="#B128-coatings-14-00822" class="html-bibr">128</a>].</p>
Full article ">Figure 16
<p>SEM images of the counterfaces used on composite coatings at (<b>a</b>) 25 °C, (<b>b</b>) 400 °C, and (<b>c</b>) 575 °C; (<b>d</b>–<b>f</b>) Raman spectra at corresponding places shown with a box. Reproduced from [<a href="#B138-coatings-14-00822" class="html-bibr">138</a>].</p>
Full article ">Figure 17
<p>Wear rate as a function of (<b>a</b>) load, (<b>b</b>) sliding velocity, and (<b>c</b>) heat treatment; and (<b>d</b>) friction coefficient of the coatings. Reproduced from [<a href="#B145-coatings-14-00822" class="html-bibr">145</a>].</p>
Full article ">Figure 18
<p>Slices from the vertical, horizontal, and transverse planes taken within the XCT reconstruction results, showing the diamond particle retainability on each slice. (<b>a</b>) 91.29%, (<b>b</b>) 82.27%, (<b>c</b>) 83.06%, (<b>d</b>) 76.22%. Reproduced from [<a href="#B154-coatings-14-00822" class="html-bibr">154</a>].</p>
Full article ">Figure 19
<p>SEM micrographs of GRCop-42 with varying contents of HR-1 produced using N2 process gas after the aging heat treatment, showing the reduction in σ phase with increasing HR-1 content: (<b>a</b>) GRCop-42-15wt% HR-1; (<b>b</b>) GRCop-42-25wt% HR-1; (<b>c</b>) GRCop-42-50wt% HR-1; (<b>d</b>) GRCop-42-75wt% HR-1; and (<b>e</b>) GRCop42-85wt% HR-1. Reproduced from [<a href="#B164-coatings-14-00822" class="html-bibr">164</a>].</p>
Full article ">
15 pages, 1750 KiB  
Article
Density Functional Theory Studies on Boron Nitride and Silicon Carbide Nanoclusters Functionalized with Amino Acids for Organophosphorus Pesticide Adsorption
by Chia Ming Chang and Yu-Hsuan Chang
Crystals 2024, 14(7), 594; https://doi.org/10.3390/cryst14070594 - 27 Jun 2024
Cited by 1 | Viewed by 713
Abstract
This study compares the properties of B12N12 and Si12C12 nanoclusters functionalized with tyrosine in the adsorption of organophosphorus pesticides, focusing on adsorption energy and electronic stability. The results indicate that B12N12/tyrosine exhibits more [...] Read more.
This study compares the properties of B12N12 and Si12C12 nanoclusters functionalized with tyrosine in the adsorption of organophosphorus pesticides, focusing on adsorption energy and electronic stability. The results indicate that B12N12/tyrosine exhibits more negative adsorption energies than Si12C12/tyrosine, suggesting stronger interactions and higher adsorption stability. Additionally, B12N12 demonstrates higher ionization energy and chemical hardness, enhancing its electronic stability during the adsorption process. In contrast, Si12C12 has higher electrophilicity and maximum electron transfer capacity, leading to greater variability in adsorption energy and more flexible electronic structure adjustments. These findings suggest that B12N12 nanoclusters have greater potential and application value as adsorption materials, particularly when modified with tyrosine. B12N12/tyrosine demonstrates higher stability and predictability in pesticide adsorption, making it more suitable for related applications. Full article
(This article belongs to the Section Hybrid and Composite Crystalline Materials)
Show Figures

Figure 1

Figure 1
<p>Tyr124 and Tyr337 are common amino acid sites where acetylcholinesterase (AChE) forms hydrogen bonds with organophosphorus (OP) pesticides ((<b>a</b>) and (<b>b</b>)).</p>
Full article ">Figure 2
<p>After docking AChE with the OPs (Par, Cou, Dia, and Nal), the structures of Tyr337_OP were extracted and used as the starting structures for modification with B<sub>12</sub>N<sub>12</sub> nanoclusters. These modified structures were then subjected to DFT geometry optimization.</p>
Full article ">Figure 3
<p>After docking AChE with the OPs (Tri, Sul, Isa, and Chl), the structures of Tyr337_OP were extracted and used as the starting structures for modification with B<sub>12</sub>N<sub>12</sub> nanoclusters. These modified structures were then subjected to DFT geometry optimization.</p>
Full article ">Figure 4
<p>Linear correlation between adsorption energies (ΔE (kcal/mol)) and HSAB parameters (A (eV), electron affinity; χ (eV), electronegativity; η (eV), chemical hardness; S (eV<sup>−1</sup>), chemical softness; ω (eV), electrophilicity index; and ΔN<sub>max</sub>, the maximum amount of electronic charge transfer) (blue circle: B<sub>12</sub>N<sub>12</sub>_Tyr337; red square: Si<sub>12</sub>C<sub>12</sub>_Tyr337).</p>
Full article ">Figure 5
<p>The ΔGAP (eV) values after the adsorption of eight OP pesticides (blue bar: B<sub>12</sub>N<sub>12</sub>_Tyr337; red bar: Si<sub>12</sub>C<sub>12</sub>_Tyr337).</p>
Full article ">
15 pages, 2571 KiB  
Article
Nitrogen-Related High-Spin Vacancy Defects in Bulk (SiC) and 2D (hBN) Crystals: Comparative Magnetic Resonance (EPR and ENDOR) Study
by Larisa Latypova, Fadis Murzakhanov, George Mamin, Margarita Sadovnikova, Hans Jurgen von Bardeleben and Marat Gafurov
Quantum Rep. 2024, 6(2), 263-277; https://doi.org/10.3390/quantum6020019 - 14 Jun 2024
Viewed by 1523
Abstract
The distinct spin, optical, and coherence characteristics of solid-state spin defects in semiconductors have positioned them as potential qubits for quantum technologies. Both bulk and two-dimensional materials, with varying structural properties, can serve as crystalline hosts for color centers. In this study, we [...] Read more.
The distinct spin, optical, and coherence characteristics of solid-state spin defects in semiconductors have positioned them as potential qubits for quantum technologies. Both bulk and two-dimensional materials, with varying structural properties, can serve as crystalline hosts for color centers. In this study, we conduct a comparative analysis of the spin–optical, electron–nuclear, and relaxation properties of nitrogen-bound vacancy defects using electron paramagnetic resonance (EPR) and electron–nuclear double resonance (ENDOR) techniques. We examine key parameters of the spin Hamiltonian for the nitrogen vacancy (NV) center in 4H-SiC: D = 1.3 GHz, Azz = 1.1 MHz, and CQ = 2.53 MHz, as well as for the boron vacancy (VB) in hBN: D = 3.6 GHz, Azz = 85 MHz, and CQ = 2.11 MHz, and their dependence on the material matrix. The spin–spin relaxation times T2 (NV center: 50 µs and VB: 15 µs) are influenced by the local nuclear environment and spin diffusion while Rabi oscillation damping times depend on crystal size and the spatial distribution of microwave excitation. The ENDOR absorption width varies significantly among color centers due to differences in crystal structures. These findings underscore the importance of selecting an appropriate material platform for developing quantum registers based on high-spin color centers in quantum information systems. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) hBN crystals mounted on an aluminum substrate before electron irradiation. The distance between the black horizontal lines on the right is 5 mm; (<b>b</b>) Samples under study prepared for high-frequency part of the spectrometer. The characteristic dimensions of the samples and capillaries correspond to the internal diameter of the resonator to achieve the highest filling factor; (<b>c</b>) Bulk crystal (0.42 × 0.67 × 1.22 mm<sup>3</sup>) of silicon carbide under an optical microscope during the preparation of samples for experiments; (<b>d</b>) Bruker Elexsys E680 spectrometer operating at 94 GHz (W-band) equipped with helium flow cryostat; (<b>e</b>) Measurement setup diagram including the main blocks of the spectrometer for the photoinduced EPR and ENDOR.</p>
Full article ">Figure 2
<p>(<b>a</b>) ESE-EPR spectra for an <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in 4H-SiC (top half, red line) and a <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN (bottom half, green line—experiment; blue solid line—simulation). The two insets at top show the detailed recorded low- and high-field components (red solid lines at 532 nm and navy color—“dark” mode) for structurally nonequivalent centers along with the corresponding simulation (blue dashed line). Yellow arrows indicate splittings between the components of the “zero-field splitting”; an asterisk (hBN) and a dot (SiC) indicate optically neutral signals both with spin = 1/2 from ionic compensators and interstitial defects, respectively, and are outside the scope of our study. (<b>b</b>) Schematic of spin polarization of color centers under optical excitation, where GS is a ground state, ES is an excited state, and MS is a metastable state. <span class="html-italic">D</span> denotes zero-field splitting.</p>
Full article ">Figure 3
<p>Dynamic characteristics of color centers obtained at <span class="html-italic">Temp.</span> = 10 K and optical excitation with λ = 532 nm. The upper part shows the curves of Rabi oscillations (blue dots) and transverse relaxation time (red solid line in the inset) for <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> centers in SiC, and for <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN (Rabi oscillations are shown as green dots, transverse relaxations are shown as a solid dark green line). Red dashed lines for each center show decay traces of Rabi oscillations with characteristic damping time <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>τ</mi> </mrow> <mrow> <mi mathvariant="normal">R</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) EPR spectra of color centers at <span class="html-italic">Temp.</span> = 297 K, where the green solid line is a <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN, the red line in the inset is an <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in SiC. The middle peak marked by a violet asterisk on the inset refers to an interstitial defect with electron spin <span class="html-italic">S</span> = ½. This spin center is independent of optical excitation of any wavelength (260–980 nm) and is beyond the scope of our study. (<b>b</b>) Spin–spin (<span class="html-italic">T</span><sub>2</sub>) or transverse relaxation and spin–lattice (<span class="html-italic">T</span><sub>1</sub>) or longitudinal relaxation (inset) curves for both color centers, where green is the <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>V</mi> </mrow> <mrow> <mi>B</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msubsup> </mrow> </semantics></math> in hBN, red is the <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>N</mi> <mi>V</mi> </mrow> <mrow> <mo>−</mo> </mrow> </msup> </mrow> </semantics></math> center in SiC.</p>
Full article ">Figure 5
<p>ENDOR spectra for SiC and hBN irradiated crystals. Hyperfine and quadrupole splitting values of the spin Hamiltonian (1) are shown in <a href="#quantumrep-06-00019-t004" class="html-table">Table 4</a>. The top inset shows individual NMR absorption lines for <sup>14</sup>N nuclei in the hBN and SiC crystal with significantly different line widths Δν.</p>
Full article ">
17 pages, 8341 KiB  
Article
Synergistic Effect of B4C and Multi-Walled CNT on Enhancing the Tribological Performance of Aluminum A383 Hybrid Composites
by Priyaranjan Samal, Himanshu Raj, Arabinda Meher, B. Surekha, Pandu R. Vundavilli and Priyaranjan Sharma
Lubricants 2024, 12(6), 213; https://doi.org/10.3390/lubricants12060213 - 11 Jun 2024
Viewed by 945
Abstract
The requirement for high-performance and energy-saving materials motivated the researchers to develop novel composite materials. This investigation focuses on utilizing aluminum alloy (A383) as the matrix material to produce hybrid metal matrix composites (HMMCs) incorporating boron carbide (B4C) and multi-walled carbon [...] Read more.
The requirement for high-performance and energy-saving materials motivated the researchers to develop novel composite materials. This investigation focuses on utilizing aluminum alloy (A383) as the matrix material to produce hybrid metal matrix composites (HMMCs) incorporating boron carbide (B4C) and multi-walled carbon nanotube (MWCNT) through a cost-effective stir casting technique. The synthesis of HMMCs involved varying the weight fractions of B4C (2%, 4%, and 6%) and MWCNT (0.5%, 1%, and 1.5%). The metallographic study was carried out by field emission scanning electron microscopy (FESEM) mapped with EDS analysis. The results indicated a uniform dispersion and robust interfacial interaction between aluminum and the reinforced particles, significantly enhancing the mechanical properties. Micro-hardness and wear characteristics of the fabricated HMMCs were investigated using Vickers microhardness testing and the pin-on-disc tribometer setup. The disc is made of hardened chromium alloy EN 31 steel of hardness 62 HRC. The applied load was varied as 10N, 20N, 30N with a constant sliding speed of 1.5 m/s for different sliding distances. The micro-hardness value of composites reinforced with 1.5 wt% MWCNT and 6 wt% B4C improved by 61% compared to the base alloy. Additionally, the wear resistance of the composite material improved with increasing reinforcement content. Incorporating 1.5% CNT and 6% B4C as reinforcements results in the composite experiencing about a 40% reduction in wear loss compared to the unreinforced aluminum alloy matrix. Furthermore, the volumetric wear loss of the HMMCs was critically analyzed with respect to different applied loads and sliding distances. This research underscores the positive impact of varying the reinforcement content on the mechanical and wear properties of aluminum alloy-based hybrid metal matrix composites. Full article
Show Figures

Figure 1

Figure 1
<p>FESEM micrographs of (<b>a</b>) B<sub>4</sub>C powder particles, (<b>b</b>) MWCNT powder particles.</p>
Full article ">Figure 2
<p>Schematic diagram of the experimental setup for stir casting process.</p>
Full article ">Figure 3
<p>Process flow chart for the present research.</p>
Full article ">Figure 4
<p>FESEM and EDX micrographs of A383 hybrid composites with (<b>a</b>) 2% B<sub>4</sub>C-1% CNT, (<b>b</b>) 4% B<sub>4</sub>C-1% CNT, (<b>c</b>) 6% B<sub>4</sub>C-1% CNT, and (<b>d</b>) higher magnification showing CNT cluster.</p>
Full article ">Figure 5
<p>EDS Elemental analysis of cast aluminum alloy MMCs (Al + 0.5% MWCNT + 4% B<sub>4</sub>C).</p>
Full article ">Figure 6
<p>EDS Elemental analysis of cast aluminum alloy MMCs (Al + 1.5% MWCNT + 6% B<sub>4</sub>C).</p>
Full article ">Figure 7
<p>Variation of microhardness with weight percentage of reinforcements: (<b>a</b>) 0.5% CNT, (<b>b</b>) 1% CNT, (<b>c</b>) 1.5% CNT.</p>
Full article ">Figure 7 Cont.
<p>Variation of microhardness with weight percentage of reinforcements: (<b>a</b>) 0.5% CNT, (<b>b</b>) 1% CNT, (<b>c</b>) 1.5% CNT.</p>
Full article ">Figure 8
<p>Variation of volumetric wear loss with sliding distance at an applied load 10 N with (<b>a</b>) 0.5 wt% CNT, (<b>b</b>) 1 wt% CNT, (<b>c</b>) 1.5 wt% CNT.</p>
Full article ">Figure 9
<p>Protective oxide layers as shown in the SEM micrograph.</p>
Full article ">Figure 10
<p>Variation of volumetric wear loss with applied load at 1000 m sliding distance at different compositions: (<b>a</b>) 0.5 wt% CNT, (<b>b</b>) 1 wt% CNT, (<b>c</b>) 1.5 wt% CNT.</p>
Full article ">Figure 11
<p>Wear surface morphology of Al-6%B<sub>4</sub>C composites with (<b>a</b>) 0.5 wt% CNT, (<b>b</b>) 1 wt% CNT, (<b>c</b>) 1.5 wt% CNT.</p>
Full article ">Figure 12
<p>Morphology of wear debris generated during the test.</p>
Full article ">
13 pages, 6421 KiB  
Article
The Microstructure, Mechanical, and Friction-Wear Properties of Boron Carbide-Based Composites with TiB2 and SiC Formed In Situ by Reactive Spark Plasma Sintering
by Agnieszka Twardowska and Marcin Kowalski
Materials 2024, 17(10), 2379; https://doi.org/10.3390/ma17102379 - 16 May 2024
Viewed by 811
Abstract
The paper presents the influence of the temperature of the sintering process on the microstructure and selected properties of boron carbide/TiB2/SiC composites obtained in situ by spark plasma sintering (SPS). The homogeneous mixture of boron carbide and 5% vol. Ti5 [...] Read more.
The paper presents the influence of the temperature of the sintering process on the microstructure and selected properties of boron carbide/TiB2/SiC composites obtained in situ by spark plasma sintering (SPS). The homogeneous mixture of boron carbide and 5% vol. Ti5Si3 micropowders were used as the initial material. Spark plasma sintering was conducted at 1700 °C, 1800 °C, and 1900 °C for 10 min after the initial pressing at 35 MPa. The heating and cooling rate was 200 °C/min. The obtained boron carbide composites were subjected to density measurement, an analysis of the chemical and phase composition, microstructure examination, and dry friction-wear tests in ball-on-disc geometry using WC as a counterpart material. The phase compositions of the produced composites differed from the composition of the initial powder mixture. Instead of titanium silicide, two new phases appeared: TiB2 and SiC. The complete disappearance of Ti5Si3 was accompanied by a decrease in the boron carbide content of the stoichiometry formula B13C2 and an increase in the content of TiB2, while the SiC content was almost constant. The relative density of the obtained boron carbide composites, as well as their hardness and resistance to wear, increased with the sintering temperature and TiB2 content. Unfortunately, the reactions occurring during sintering did not allow us to obtain composites with high density and hardness. The relative density was 76–85.2% of the theoretical one, while the Vickers hardness was in the range of 4–12 GPa. The mechanism wear of boron carbide composites tested in friction contact with WC was abrasive. The volumetric wear rate (Wv) of composites decreased with increasing sintering temperature and TiB2 content. The average value of coefficient of friction (CoF) was in the range of 0.54–0.61, i.e., it did not differ significantly from the value for B4C sinters. Full article
Show Figures

Figure 1

Figure 1
<p>B<sub>4</sub>C/5% vol. Ti<sub>5</sub>Si<sub>3</sub> powder mixture after homogenization: (<b>a</b>) SEM image; (<b>b</b>) EDS spectrum taken from the presented area; (<b>c</b>) particle size distribution.</p>
Full article ">Figure 2
<p>(<b>a</b>) Schematic representation of an SPS sintering device. (<b>b</b>) Components of the sintering die.</p>
Full article ">Figure 3
<p>XRD patterns registered in BB geometry for composites obtained from B<sub>4</sub>C/5% vol. Ti<sub>5</sub>Si<sub>3</sub> powder with indexed peak positions for identified phases.</p>
Full article ">Figure 4
<p>SEM image of the microstructure of B<sub>4</sub>C-TiB<sub>2</sub>-SiC composites with EDS elemental maps of Ti, Si, and B distribution: (<b>a</b>) 1800 °C; (<b>b</b>) 1900 °C.</p>
Full article ">Figure 5
<p>(<b>a</b>) The microstructure (SEM image) of the composite sintered at 1900 °C with (<b>b</b>) EDS spectra taken from marked area (spectrum 1–3) and whole area (spectrum 4).</p>
Full article ">Figure 6
<p>SEM images of the surface of the composite sintered at 1900 °C after hardness measurement in indented area: (<b>a</b>) secondary electron image; (<b>b</b>) backscattered electron image in topography mode (BET).</p>
Full article ">Figure 7
<p>Exemplary results of roughness and volume loss measurement of produced composites by confocal microscopy after ball-on-disc test (in dry sliding with WC ball as a counterpart material): (<b>a</b>) surface height 3D map; (<b>b</b>) roughness profile; (<b>c</b>) surface area of wear track cross-section for volume loss measurement.</p>
Full article ">Figure 8
<p>Experimental plots of CoF and wear rate of produced composites, sintered at (<b>a</b>) 1800 °C; (<b>b</b>) 1900 °C.</p>
Full article ">Figure 9
<p>CoF values and volumetric wear indexes (Wv) for boron carbide composites determined in friction-wear test with WC as counterpart material.</p>
Full article ">Figure 10
<p>SEM image of the surface of B<sub>4</sub>C composite sintered at 1900 °C after friction-wear test (WC as counterpart material), with EDS maps showing distribution of B, Ti, Si, W, and O.%.</p>
Full article ">
13 pages, 6821 KiB  
Article
Synergistic Effects of Boron and Rare Earth Elements on the Microstructure and Stress Rupture Properties in a Ni-Based Superalloy
by Qiang Tian, Shuo Huang, Heyong Qin, Ran Duan, Chong Wang and Xintong Lian
Materials 2024, 17(9), 2007; https://doi.org/10.3390/ma17092007 - 25 Apr 2024
Viewed by 836
Abstract
The synergistic effects of boron (B) and rare earth (RE) elements on the microstructure and stress rupture properties were investigated in a Ni-based superalloy. The stress rupture lifetime at 650 °C/873 MPa significantly increased with the addition of B as a single element. [...] Read more.
The synergistic effects of boron (B) and rare earth (RE) elements on the microstructure and stress rupture properties were investigated in a Ni-based superalloy. The stress rupture lifetime at 650 °C/873 MPa significantly increased with the addition of B as a single element. Furthermore, the stress rupture lifetime reached its peak (303 h), with a certain amount of B and RE added together in test alloys. Although the grain size and morphology of the γ′ phase varied a little with the change in B and RE addition, they were not considered to be the main reasons for stress rupture performance. The enhancement in stress rupture lifetime was mostly attributed to the segregation of the B and RE elements, which increased the binding force of the grain boundary and improved its strength and plasticity. In addition, the enrichment of B and RE inhabited the precipitation of carbides along grain boundaries. Furthermore, nano-scale RE precipitates containing sulfur (S) and phosphorus (P) were observed to be distributed along the grain boundaries. The purification of grain boundaries by B and RE elements was favorable to further improve the stress rupture properties. Full article
(This article belongs to the Section Advanced Materials Characterization)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram for stress rupture specimens (mm).</p>
Full article ">Figure 2
<p>Microstructure of different GH4742 test alloys: (<b>a</b>) alloy 1; (<b>b</b>) alloy 2; (<b>c</b>) alloy 3; (<b>d</b>) alloy 4; (<b>e</b>) alloy 5.</p>
Full article ">Figure 3
<p>SEM images of γ′ phase of different GH4742 test alloys: (<b>a</b>) alloy 1; (<b>b</b>) alloy 2; (<b>c</b>) alloy 3; (<b>d</b>) alloy 4; (<b>e</b>) alloy 5.</p>
Full article ">Figure 4
<p>EPMA images of element distribution in precipitate towards grain boundary in alloy 5.</p>
Full article ">Figure 5
<p>Stress rupture properties of different GH4742 test alloys.</p>
Full article ">Figure 6
<p>Morphologies of fracture surface of test alloys: (<b>a</b>,<b>f</b>,<b>k</b>) alloy 1; (<b>b</b>,<b>g</b>,<b>l</b>) alloy 2; (<b>c</b>,<b>h</b>,<b>m</b>) alloy 3; (<b>d</b>,<b>i</b>,<b>n</b>) alloy 4; (<b>e</b>,<b>j</b>,<b>o</b>) alloy 5.</p>
Full article ">Figure 7
<p>A schematic diagram of selected areas (<b>a</b>) and its AES spectrum (<b>b</b>) in alloy 5.</p>
Full article ">Figure 8
<p>Intergranular fracture morphology (<b>a</b>) and AES spectrum of La, Ce and B (<b>b</b>) in alloy 5.</p>
Full article ">Figure 9
<p>STEM-HAADF images and elemental mapping of grain boundary precipitates in alloy 5 (<b>a</b>) RE compound with Ce, O and Cr (<b>b</b>) RE complex inclusion with C, B, La, P and S.</p>
Full article ">
Back to TopTop