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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,100)

Search Parameters:
Keywords = high entropy alloys

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
8 pages, 2738 KiB  
Communication
Predictions of Lattice Parameters in NiTi High-Entropy Shape-Memory Alloys Using Different Machine Learning Models
by Tu-Ngoc Lam, Jiajun Jiang, Min-Cheng Hsu, Shr-Ruei Tsai, Mao-Yuan Luo, Shuo-Ting Hsu, Wen-Jay Lee, Chung-Hao Chen and E-Wen Huang
Materials 2024, 17(19), 4754; https://doi.org/10.3390/ma17194754 - 27 Sep 2024
Viewed by 201
Abstract
This work applied three machine learning (ML) models—linear regression (LR), random forest (RF), and support vector regression (SVR)—to predict the lattice parameters of the monoclinic B19′ phase in two distinct training datasets: previously published ZrO2-based shape-memory ceramics (SMCs) and NiTi-based high-entropy [...] Read more.
This work applied three machine learning (ML) models—linear regression (LR), random forest (RF), and support vector regression (SVR)—to predict the lattice parameters of the monoclinic B19′ phase in two distinct training datasets: previously published ZrO2-based shape-memory ceramics (SMCs) and NiTi-based high-entropy shape-memory alloys (HESMAs). Our findings showed that LR provided the most accurate predictions for ac, am, bm, and cm in NiTi-based HESMAs, while RF excelled in computing βm for both datasets. SVR disclosed the largest deviation between the predicted and actual values of lattice parameters for both training datasets. A combination approach of RF and LR models enhanced the accuracy of predicting lattice parameters of martensitic phases in various shape-memory materials for stable high-temperature applications. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart for the computational prediction of lattice constants in the NiTi-based HESMAs.</p>
Full article ">Figure 2
<p>The predicted versus actual lattice constants of a<sub>m</sub> (<b>a</b>), b<sub>m</sub> (<b>b</b>), c<sub>m</sub> (<b>c</b>), and β<sub>m</sub> (<b>d</b>) modeled by LR. Those of a<sub>m</sub> (<b>e</b>), b<sub>m</sub> (<b>f</b>), c<sub>m</sub> (<b>g</b>), and β<sub>m</sub> (<b>h</b>) modeled by RF. Those of a<sub>m</sub> (<b>i</b>), b<sub>m</sub> (<b>j</b>), c<sub>m</sub> (<b>k</b>), and β<sub>m</sub> (<b>l</b>) modeled by SVR in the ZrO<sub>2</sub>-based SMCs. The solid black line depicts the perfect match between the predicted versus actual values of lattice constants.</p>
Full article ">Figure 3
<p>The predicted versus actual lattice constants of a<sub>m</sub> (<b>a</b>), b<sub>m</sub> (<b>b</b>), c<sub>m</sub> (<b>c</b>), β<sub>m</sub> (<b>d</b>), and a<sub>c</sub> (<b>e</b>) modeled by LR. Those of a<sub>m</sub> (<b>f</b>), b<sub>m</sub> (<b>g</b>), c<sub>m</sub> (<b>h</b>), β<sub>m</sub> (<b>i</b>), and a<sub>c</sub> (<b>j</b>) modeled by RF. Those of a<sub>m</sub> (<b>k</b>), b<sub>m</sub> (<b>l</b>), c<sub>m</sub> (<b>m</b>), β<sub>m</sub> (<b>n</b>), and a<sub>c</sub> (<b>o</b>) modeled by SVR in the NiTi-based HESMAs. The solid black line depicts the perfect match between the predicted versus actual values of lattice constants.</p>
Full article ">Figure 4
<p>The test RMSE values among three ML models in computing the predicted lattice parameters of a<sub>m</sub>, b<sub>m</sub>, c<sub>m</sub>, and β<sub>m</sub> in the ZrO<sub>2</sub>-based SMCs (<b>a</b>). Those of a<sub>c</sub>, a<sub>m</sub>, b<sub>m</sub>, c<sub>m</sub>, and β<sub>m</sub> in the NiTi-based HESMAs (<b>b</b>).</p>
Full article ">
15 pages, 7008 KiB  
Article
Radiation Resistance of High-Entropy Alloys CoCrFeNi and CoCrFeMnNi, Sequentially Irradiated with Kr and He Ions
by Bauyrzhan Amanzhulov, Igor Ivanov, Vladimir Uglov, Sergey Zlotski, Azamat Ryskulov, Alisher Kurakhmedov, Asset Sapar, Yerulan Ungarbayev, Mikhail Koloberdin and Maxim Zdorovets
Materials 2024, 17(19), 4751; https://doi.org/10.3390/ma17194751 - 27 Sep 2024
Viewed by 202
Abstract
This work studied the effect of sequential irradiation by krypton and helium ions at room temperature on the composition and structure of CoCrFeNi and CoCrFeMnNi high-entropy alloys (HEAs). Irradiation of the HEAs by 280 keV Kr14+ ions up to a fluence of [...] Read more.
This work studied the effect of sequential irradiation by krypton and helium ions at room temperature on the composition and structure of CoCrFeNi and CoCrFeMnNi high-entropy alloys (HEAs). Irradiation of the HEAs by 280 keV Kr14+ ions up to a fluence of 5 × 1015 cm–2 and 40 keV He2+ ions up to a fluence of 2 × 1017 cm–2 did not alter their elemental distribution and constituent phases. Blisters formed on the nickel surface after sequential irradiation, where large blisters had an average diameter of 3.8 μm. The lattice parameter of the (Co, Cr, Fe and Ni) and (Co, Cr, Fe, Mn and Ni) solid solutions increased by 0.17% and 0.37% after sequential irradiation, respectively. Irradiation by Kr ions led to a decrease in tensile macrostresses in the HEAs in the region of krypton ion implantation (Region I) and the formation of compressive macrostresses in the region behind the peak of implanted krypton (Region II). Sequential irradiation formed large compressive stresses in Ni and HEAs equal to −131.5 MPa, −300 MPa and −613.5 MPa in Ni, CoCrFeNi and CoCrFeMnNi, respectively, in the Region II. Irradiation by krypton ions decreased the dislocation density by 1.6–2.3 times, and irradiation with helium ions increased it by 11–15 times relative to unirradiated samples for CoCrFeNi and CoCrFeMnNi, respectively. Sequentially irradiated CoCrFeMnNi HEA had higher macrostresses and dislocation density than CoCrFeNi. Full article
(This article belongs to the Special Issue Advanced Science and Technology of High Entropy Materials)
Show Figures

Figure 1

Figure 1
<p>The profile of (<b>a</b>) implanted krypton and helium and (<b>b</b>) radiation damage in Ni and CoCrFeNi and CoCrFeMnNi HEAs irradiated by Kr<sup>14+</sup> (280 keV, 5 × 10<sup>15</sup> cm<sup>−2</sup>) and He<sup>2+</sup> (40 keV, 2 × 10<sup>17</sup> cm<sup>−2</sup>).</p>
Full article ">Figure 2
<p>HIRBS spectra (black—experimental, red—RUMP) of: (<b>a</b>,<b>d</b>) initial [<a href="#B23-materials-17-04751" class="html-bibr">23</a>] and irradiated by (<b>b</b>,<b>e</b>) Krypton ions, (<b>c</b>,<b>f</b>) Kr and He ions, (<b>a</b>–<b>c</b>) CoCrFeNi and (<b>d</b>–<b>f</b>) CoCrFeMnNi HEAs.</p>
Full article ">Figure 3
<p>SEM images in the backscattering electrons mode (BSE) of the sample surface irradiated by: (<b>a</b>–<b>c</b>) Kr ions, where (<b>a</b>) Ni, (<b>b</b>) CoCrFeNi, (<b>c</b>) CoCrFeMnNi and their respective elemental maps.</p>
Full article ">Figure 4
<p>SEM images in the backscattering electrons mode (BSE) of the sample surface irradiated by: (<b>a</b>–<b>c</b>) Kr and He ions, where (<b>a</b>) Ni, (<b>b</b>) CoCrFeNi, (<b>c</b>) CoCrFeMnNi and their respective elemental maps.</p>
Full article ">Figure 5
<p>Diameters of blisters from the surface SEM analysis of (<b>a</b>) Ni and (<b>b</b>) CoCrFeMnNi irradiated by Kr and He ions, corresponding to <a href="#materials-17-04751-f004" class="html-fig">Figure 4</a>a and <a href="#materials-17-04751-f004" class="html-fig">Figure 4</a>c, respectively.</p>
Full article ">Figure 6
<p>XRD patterns of initial, irradiated by Kr and sequentially irradiated by Kr and He ions samples of: (<b>a</b>) Ni, (<b>b</b>) CoCrFeNi and (<b>c</b>) CoCrFeMnNi HEAs, collected at the X-ray incidence angle α.</p>
Full article ">Figure 7
<p>Macrostresses in the initial [<a href="#B23-materials-17-04751" class="html-bibr">23</a>], irradiated by Kr ions, irradiated by He ions [<a href="#B23-materials-17-04751" class="html-bibr">23</a>] and sequentially irradiated by Kr and He ions Ni, CoCrFeNi and CoCrFeMnNi HEAs at XRD scan ranges of: (<b>a</b>) 100 nm and (<b>b</b>) 300 nm.</p>
Full article ">Figure 8
<p>Relative density of dislocations in the initial [<a href="#B23-materials-17-04751" class="html-bibr">23</a>], Kr-irradiated, He-irradiated [<a href="#B23-materials-17-04751" class="html-bibr">23</a>] and sequentially irradiated samples of Ni, CoCrFeNi and CoCrFeMnNi HEAs at an XRD scan range of (<b>a</b>) 100 nm and (<b>b</b>) 300 nm.</p>
Full article ">
18 pages, 12338 KiB  
Article
Effects of Mo Addition on Microstructure and Corrosion Resistance of Cr25-xCo25Ni25Fe25Mox High-Entropy Alloys via Directed Energy Deposition
by Han-Eol Kim, Jae-Hyun Kim, Ho-In Jeong, Young-Tae Cho, Osama Salem, Dong-Won Jung and Choon-Man Lee
Micromachines 2024, 15(10), 1196; https://doi.org/10.3390/mi15101196 - 27 Sep 2024
Viewed by 262
Abstract
Highly entropy alloys (HEAs) are novel materials that have great potential for application in aerospace and marine engineering due to their superior mechanical properties and benefits over conventional materials. NiCrCoFe, also referred to as Ni-based HEA, has exceptional low-temperature strength and microstructural stability. [...] Read more.
Highly entropy alloys (HEAs) are novel materials that have great potential for application in aerospace and marine engineering due to their superior mechanical properties and benefits over conventional materials. NiCrCoFe, also referred to as Ni-based HEA, has exceptional low-temperature strength and microstructural stability. However, HEAs have limited corrosion resistance in some environments, such as a 3.5 wt% sodium chloride (NaCl) solution. Adding corrosion-resistant elements such as molybdenum (Mo) to HEAs is expected to increase their corrosion resistance in a variety of corrosive environments. Metal additive manufacturing reduces production times compared to casting and eliminates shrinkage issues, making it ideal for producing homogeneous HEA. This study used directed energy deposition (DED) to create Cr25-xCo25Ni25Fe25Mox (x = 0, 5, 10%) HEAs. Tensile strength and potentiodynamic polarization tests were used to assess the materials’ mechanical properties and corrosion resistance. The mechanical tests revealed that adding 5% Mo increased yield strength (YS) by 20.1% and ultimate tensile strength (UTS) by 9.5% when compared to 0% Mo. Adding 10% Mo led to a 32.5% increase in YS and a 20.4% increase in UTS. Potentiodynamic polarization tests were used to assess corrosion resistance in a 3.5-weight percent NaCl solution. The results showed that adding Mo significantly increased initial corrosion resistance. The alloy with 5% Mo had a higher corrosion potential (Ecorr) and a lower current density (Icorr) than the alloy with 0% Mo, indicating improved initial corrosion resistance. The alloy containing 10% Mo had the highest corrosion potential and the lowest current density, indicating the slowest corrosion rate and the best initial corrosion resistance. Finally, Cr25-xCo25Ni25Fe25Mox (x = 0, 5, 10%) HEAs produced by DED exhibited excellent mechanical properties and corrosion resistance, which can be attributed to the presence of Mo. Full article
(This article belongs to the Special Issue Future Prospects of Additive Manufacturing)
Show Figures

Figure 1

Figure 1
<p>The schematic diagram for the DED process [<a href="#B19-micromachines-15-01196" class="html-bibr">19</a>].</p>
Full article ">Figure 2
<p>Schematic illustration of the overall research work.</p>
Full article ">Figure 3
<p>The experimental setup for the DED process.</p>
Full article ">Figure 4
<p>The SEM images of Cr, Co, Ni, Fe, and Mo powder.</p>
Full article ">Figure 5
<p>The initial experimental results about the hatching distance.</p>
Full article ">Figure 6
<p>The extracted sampling positions for performing analysis.</p>
Full article ">Figure 7
<p>The XRD results of the Cr<sub>25-x</sub>Co<sub>25</sub>Ni<sub>25</sub>Fe<sub>25</sub>Mo<sub>x</sub> (x = 0, 5, 10) specimen.</p>
Full article ">Figure 8
<p>EBSD IPF mapping and grain size of (<b>a</b>) 0%, (<b>b</b>) 5%, and (<b>c</b>) 10% Mo specimen.</p>
Full article ">Figure 9
<p>SEM image of (<b>a</b>) 0%, (<b>b</b>) 5%, and (<b>c</b>) 10% Mo specimen.</p>
Full article ">Figure 10
<p>EDS results and a table analyzing the composition of each point of the (<b>a</b>) 0%, (<b>b</b>) 5%, and (<b>c</b>) 10% Mo specimen.</p>
Full article ">Figure 11
<p>Tensile test results of Cr<sub>25-x</sub>Co<sub>25</sub>Ni<sub>25</sub>Fe<sub>25</sub>Mo<sub>x</sub> (at x = 0, 5, 10%) specimen.</p>
Full article ">Figure 12
<p>SEM image of fracture surfaces of (<b>a</b>) 0%, (<b>b</b>) 5%, and (<b>c</b>) 10% Mo specimen.</p>
Full article ">Figure 13
<p>The micrograph of the specimen results of corrosion tests.</p>
Full article ">Figure 14
<p>Potentiodynamic polarization test curves of Cr<sub>25-x</sub>Co<sub>25</sub>Ni<sub>25</sub>Fe<sub>25</sub>Mo<sub>x</sub> (at x = 0, 5, 10%) specimen.</p>
Full article ">
19 pages, 3747 KiB  
Article
Ductility Index for Refractory High Entropy Alloys
by Ottó K. Temesi, Lajos K. Varga, Nguyen Quang Chinh and Levente Vitos
Crystals 2024, 14(10), 838; https://doi.org/10.3390/cryst14100838 - 27 Sep 2024
Viewed by 160
Abstract
The big advantage of refractory high entropy alloys (RHEAs) is their strength at high temperatures, but their big disadvantage is their brittleness at room temperature, which prevents their machining. There is a great need to classify the alloys in terms of brittle-ductile (B-D) [...] Read more.
The big advantage of refractory high entropy alloys (RHEAs) is their strength at high temperatures, but their big disadvantage is their brittleness at room temperature, which prevents their machining. There is a great need to classify the alloys in terms of brittle-ductile (B-D) properties, with easily obtainable ductility indices (DIs) ready to help design these refractory alloys. Usually, the DIs are checked by representing them as a function of fraction strain, ε. The critical values of DI and ε divide the DI—ε area into four squares. In the case of a successful DI, the points representing the alloys are located in the two diagonal opposite squares, well separating the alloys with (B-D) properties. However, due to the scatter of the data, the B-D separation is not perfect, and it is difficult to establish the critical value of DI. In this paper, we solve this problem by replacing the fracture strain parameter with new DIs that scale with the old DIs. These new DIs are based on the force constant and amplitude of thermal vibration around the Debye temperature. All of them are easily available and can be calculated from tabulated data. Full article
(This article belongs to the Special Issue Advances in Processing, Simulation and Characterization of Alloys)
Show Figures

Figure 1

Figure 1
<p>Fracture energy versus fracture strain. Two opposite quadrants separately contain the brittle and ductile alloys, B and D. The other two quarters are empty if the critical values are chosen correctly. ε<sub>c</sub> = 10% and E<sub>c</sub> = 200 MJ/m<sup>3</sup>.</p>
Full article ">Figure 2
<p>Fracture strain ε as a function of reduced VEC. The critical values are ε<sub>c</sub> = 10%, R-VEC<sub>c</sub> = 0.72, and VEC<sub>cr</sub> = 5.4.</p>
Full article ">Figure 3
<p>Fracture strain c as a function of modified Poisson ratio parameter D*. The critical values are ε<sub>c</sub> = 10% and D<sub>c</sub>* = 0.562.</p>
Full article ">Figure 4
<p>The D parameter of Rice scales with the D* of Christensen.</p>
Full article ">Figure 5
<p>Fracture strain ε as a function of the D parameter. The critical values are ε<sub>c</sub> = 10% and D<sub>c</sub> = 2.51.</p>
Full article ">Figure 6
<p>Valence electron counts dependent of the C′ component of the shear modulus for refractory metals and their alloys. Linear correlation was found between the valence electron counts and the C′ component of shear modulus for refractory metals and alloys. For alloys the data from [<a href="#B18-crystals-14-00838" class="html-bibr">18</a>,<a href="#B30-crystals-14-00838" class="html-bibr">30</a>] and for pure metals the tabulated data have been used.</p>
Full article ">Figure 7
<p>Delimitation of the B-D regions using Cauchy pressure parameter P = 1 − 2.66<sup>x</sup> (<span class="html-italic">G/B</span>) versus (<b>a</b>) fracture strain and (<b>b</b>) <span class="html-italic">R-VEC</span>.</p>
Full article ">Figure 8
<p>Cauchy pressure <span class="html-italic">C</span>″ = PSM = 1 − 2.66 G/B as a function of the <span class="html-italic">D</span> parameter. Again, we obtained the critical value <span class="html-italic">D<sub>c</sub></span> = 2.5.</p>
Full article ">Figure 9
<p>Delimitation of B-D regions using Cauchy pressure, P = 1 − 2.66<sup>x</sup><span class="html-italic">G/B</span>), and the Pugh ratio versus <span class="html-italic">D</span> parameter.</p>
Full article ">Figure 10
<p>Scaling of the shear moduli with the force constant for elemental metals (<b>a</b>) and alloys (<b>b</b>).</p>
Full article ">Figure 11
<p>Scaling of the shear moduli with the ratio of the force constant and thermal amplitude for elemental metals.</p>
Full article ">Figure 12
<p>Scaling of the bulk modulus with the ratio of melting point and molar volume.</p>
Full article ">Figure 13
<p>Clear delimitation of D and B quadrants in G/B versus sqrt (RNP) representations for (<b>a</b>) elemental metals and (<b>b</b>) for RHEAs.</p>
Full article ">Figure 14
<p>Surface energy of metals scaled with the melting temperature, similar to the cohesion energy.</p>
Full article ">Figure 15
<p>Scaling the R-T parameter with k/γ<sub>surf,</sub> for elemental metals. The critical value (k/γ)<sub>cr</sub>~185, whereas (Gb/γ)<sub>cr</sub> = 10.</p>
Full article ">Figure 16
<p>Assignment in B-D regions in the representation of old and new R-T parameters. The alloys were taken from the [<a href="#B35-crystals-14-00838" class="html-bibr">35</a>] dataset.</p>
Full article ">
18 pages, 6623 KiB  
Article
Effect of ZrO2 Particles on the Microstructure and Ultrasonic Cavitation Properties of CoCrFeMnNi High-Entropy Alloy Composite Coatings
by Danqing Yin, Junming Chang, Yonglei Wang, Ning Ma, Junnan Zhao, Haoqi Zhao and Meng Wang
Coatings 2024, 14(10), 1235; https://doi.org/10.3390/coatings14101235 - 25 Sep 2024
Viewed by 452
Abstract
CoCrFeMnNi-XZrO2 (X is a mass percentage, X = 1, 3, 5, and 10) high-entropy alloy composite coatings were successfully prepared on 0Cr13Ni5Mo martensitic stainless steel substrates using laser cladding technology. The phase composition, microstructure, mechanical properties, and cavitation erosion behavior of the [...] Read more.
CoCrFeMnNi-XZrO2 (X is a mass percentage, X = 1, 3, 5, and 10) high-entropy alloy composite coatings were successfully prepared on 0Cr13Ni5Mo martensitic stainless steel substrates using laser cladding technology. The phase composition, microstructure, mechanical properties, and cavitation erosion behavior of the composite coatings under different contents of ZrO2 were studied. The mechanism of ZrO2 particle-reinforced cavitation corrosion resistance was studied using ABAQUS2023 finite element software. The results show that the phase structure of the composite coating organization is composed of FCC phase reinforced by ZrO2 phase. The addition of ZrO2 causes lattice distortion. The coatings have typical branch crystals and an equiaxed crystal microstructure. With the increase in ZrO2 content, the microhardness of the composite coatings gradually increases. When X = 10%, the coating’s microhardness reached 348 HV, which was 95.53% higher than the high-entropy alloys without ZrO2 added. Adding ZrO2 can prolong the incubation period of high-entropy alloys; the high-entropy alloy composite coating with 5 wt.% ZrO2 exhibited the best cavitation resistance, with a cumulative volume loss rate of only 15.74% of the substrate after 10 h of ultrasonic cavitation erosion. The simulation results indicate that ZrO2 can withstand higher stress and deformation in cavitation erosion, reduce the degree of substrate damage, and generate higher compressive stress on the coating surface to cope with cavitation erosion. Full article
Show Figures

Figure 1

Figure 1
<p>SEM images of the hybrid powders and their corresponding energy spectral surface distribution:(<b>a</b>) SEM image of the hybrid powders, (<b>b</b>) distribution of six elements (<b>c</b>) distribution of Co element, (<b>d</b>) distribution of Cr element, (<b>e</b>) distribution of Ni element, (<b>f</b>) distribution of Mn element, (<b>g</b>) distribution of Fe element, (<b>h</b>) distribution of Zr element.</p>
Full article ">Figure 2
<p>Schematic diagram of laser cladding system.</p>
Full article ">Figure 3
<p>Schematic illustration of ultrasonication set-up used for cavitation erosion.</p>
Full article ">Figure 4
<p>Schematic illustration of finite element model; (<b>a</b>) 3D model; (<b>b</b>) 2D model.</p>
Full article ">Figure 5
<p>XRD pattern and local magnification of S1–S5.</p>
Full article ">Figure 6
<p>Macroscopic morphology after laser cladding: (<b>a</b>) S1; (<b>b</b>) S2; (<b>c</b>) S3; (<b>d</b>) S4; and (<b>e</b>) S5.</p>
Full article ">Figure 7
<p>(<b>a</b>) BSE image of coatings and substrate, (<b>b</b>) Local magnification of pores.</p>
Full article ">Figure 8
<p>SEM image of HEAs: (<b>a</b>) S1; (<b>b</b>) S2; (<b>c</b>) S3; (<b>d</b>) S4; (<b>e</b>) and S5.</p>
Full article ">Figure 9
<p>Maps of elements distribution and EDS analysis of S2 by SEM area scan: (<b>a</b>) SEM image of CoCrFeMnNi-1%ZrO<sub>2</sub>; (<b>b</b>) distribution of Cr element; (<b>c</b>) distribution of Fe element; (<b>d</b>) distribution of Mn element; (<b>e</b>) distribution of Co element; and (<b>f</b>) distribution of Zr element.</p>
Full article ">Figure 10
<p>(<b>a</b>) SEM of area A in <a href="#coatings-14-01235-f009" class="html-fig">Figure 9</a>a; (<b>b</b>) distribution of Zr element; and (<b>c</b>) EDS of area A.</p>
Full article ">Figure 11
<p>(<b>a</b>) Average microhardness of the substrate and coatings. (<b>b</b>) Microhardness distribution of the coatings.</p>
Full article ">Figure 12
<p>(<b>a</b>) CE cumulative volume loss as and (<b>b</b>) MDER results for substrates S1, S2, S3, S4, and S5 after 10 h CE in the deionized water.</p>
Full article ">Figure 13
<p>Cavitation erosion morphologies of substrate and HEACs after 1 h and 10 h: (<b>a1</b>,<b>a2</b>) 0Cr13Ni5Mo; (<b>b1</b>,<b>b2</b>) S1; (<b>c1</b>,<b>c2</b>) S2; (<b>d1</b>,<b>d2</b>) S3; (<b>e1</b>,<b>e2</b>) S4; and (<b>f1</b>,<b>f2</b>) S5. (<b>a1</b>–<b>f1</b>) 1 h cavitation erosion; (<b>a2</b>–<b>f2</b>) 10 h cavitation erosion.</p>
Full article ">Figure 14
<p>Three-dimensional images of cavitation erosion morphologies of HEACs: (<b>a</b>) 0Cr13Ni5Mo; (<b>b</b>) S1; (<b>c</b>) S2; (<b>d</b>) S3; (<b>e</b>) S4; and (<b>f</b>) S5.</p>
Full article ">Figure 15
<p>(<b>a</b>) Displacement variation of S1–S5 3D model and (<b>b</b>) S1–S5 maximum deformation and average deformation.</p>
Full article ">Figure 16
<p>S22 stress distribution of 2D model: (<b>a</b>) S1; (<b>b</b>) S2; (<b>c</b>) S3; (<b>d</b>) S4; and (<b>e</b>) S5.</p>
Full article ">
18 pages, 14791 KiB  
Article
Effect of Substrate Bias on the Structure and Tribological Performance of (AlTiVCrNb)CxNy Coatings Deposited via Graphite Co-Sputtering
by Haichao Cai, Pengge Guo, Yujun Xue, Lulu Pei, Yinghao Zhang and Jun Ye
Lubricants 2024, 12(9), 325; https://doi.org/10.3390/lubricants12090325 - 23 Sep 2024
Viewed by 362
Abstract
In the existing literature, there are few studies on the effect of deposition bias on the tribological properties of carbon-doped high-entropy alloy coatings. In order to further study the effect of the deposition bias on the properties of coatings, (AlTiVCrNb)CxNy [...] Read more.
In the existing literature, there are few studies on the effect of deposition bias on the tribological properties of carbon-doped high-entropy alloy coatings. In order to further study the effect of the deposition bias on the properties of coatings, (AlTiVCrNb)CxNy coatings were deposited via unbalanced RF magnetron sputtering. The microstructure and tribological properties of carbon-doped high-entropy alloy ceramic coatings under different deposition biases were studied. The composition, morphology, crystal structure, and chemical morphology of each element of the coating were analyzed using scanning electron microscopy (SEM), X-ray diffraction (XRD), and X-ray photoelectron spectroscopy (XPS). The hardness, elastic modulus, friction, and wear properties of the coating were further characterized using a nanoindentation instrument, reciprocating sliding friction, a wear tester, and a white light interferometer. The coating density reached the optimal level when the deposition bias value was 90 V. The hardness and elastic modulus of the (AlTiVCrNb)CxNy coating increased first and then decreased with an increase in deposition bias, and the maximum hardness was 23.98 GPa. When the deposition bias was 90 V, the coating formed a good-quality carbon transfer film on the surface of the counterbody due to sp2 clusters during the friction and wear process. The average friction coefficient and wear rate of the (AlTiVCrNb)CxNy coating were the lowest, 0.185 and 1.6 × 10−7 mm3/N·m, respectively. The microstructure, mechanical properties, and tribological performance of the (AlTiVCrNb)CxNy coating were greatly affected by the change in deposition bias, and an (AlTiVCrNb)CxNy coating with excellent structure and friction properties could be prepared using graphite co-sputtering. Full article
Show Figures

Figure 1

Figure 1
<p>Surface and cross-sectional SEM images of (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating. (<b>a</b>) (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating surface morphology. (<b>b</b>) (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating surface morphology.</p>
Full article ">Figure 1 Cont.
<p>Surface and cross-sectional SEM images of (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating. (<b>a</b>) (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating surface morphology. (<b>b</b>) (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating surface morphology.</p>
Full article ">Figure 2
<p>XRD and Raman spectra of coatings with different Z<sub>c</sub> values (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub>. (<b>a</b>) XRD patterns. (<b>b</b>) Raman spectra.</p>
Full article ">Figure 2 Cont.
<p>XRD and Raman spectra of coatings with different Z<sub>c</sub> values (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub>. (<b>a</b>) XRD patterns. (<b>b</b>) Raman spectra.</p>
Full article ">Figure 3
<p>The (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating XPS Nb 3d, Al2p, Cr2p, V2p, Ti2p, and C1s spectra. (<b>a</b>) Nb3d spectrum. (<b>b</b>) Al2p spectrum. (<b>c</b>) Cr2p spectrum. (<b>d</b>) V2p spectrum. (<b>e</b>) Ti2p spectrum. (<b>f</b>) C1s spectrum.</p>
Full article ">Figure 3 Cont.
<p>The (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating XPS Nb 3d, Al2p, Cr2p, V2p, Ti2p, and C1s spectra. (<b>a</b>) Nb3d spectrum. (<b>b</b>) Al2p spectrum. (<b>c</b>) Cr2p spectrum. (<b>d</b>) V2p spectrum. (<b>e</b>) Ti2p spectrum. (<b>f</b>) C1s spectrum.</p>
Full article ">Figure 3 Cont.
<p>The (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating XPS Nb 3d, Al2p, Cr2p, V2p, Ti2p, and C1s spectra. (<b>a</b>) Nb3d spectrum. (<b>b</b>) Al2p spectrum. (<b>c</b>) Cr2p spectrum. (<b>d</b>) V2p spectrum. (<b>e</b>) Ti2p spectrum. (<b>f</b>) C1s spectrum.</p>
Full article ">Figure 4
<p>The hardness H and elastic modulus E, H/E, and H<sup>3</sup>/E<sup>2</sup> of (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coatings with different Z<sub>c</sub>. (<b>a</b>) Hardness. (<b>b</b>) H/E and H<sup>3</sup>/E<sup>2</sup>.</p>
Full article ">Figure 4 Cont.
<p>The hardness H and elastic modulus E, H/E, and H<sup>3</sup>/E<sup>2</sup> of (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coatings with different Z<sub>c</sub>. (<b>a</b>) Hardness. (<b>b</b>) H/E and H<sup>3</sup>/E<sup>2</sup>.</p>
Full article ">Figure 5
<p>The friction coefficient, the morphology of wear marks, and the distribution of oxygen elements in wear marks of coatings under different De values (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub>. (<b>a</b>) The friction coefficient. (<b>b</b>) The wear scar appearance. (<b>c</b>) The distribution of oxygen element in wear scar.</p>
Full article ">Figure 5 Cont.
<p>The friction coefficient, the morphology of wear marks, and the distribution of oxygen elements in wear marks of coatings under different De values (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub>. (<b>a</b>) The friction coefficient. (<b>b</b>) The wear scar appearance. (<b>c</b>) The distribution of oxygen element in wear scar.</p>
Full article ">Figure 6
<p>SEM micrographs and EDS mapping images of the wear scars on the steel balls.</p>
Full article ">Figure 7
<p>The Raman spectrum of the coated samples. (<b>a</b>) The Raman spectra of the wear tracks. (<b>b</b>) The Raman spectra of the wear scars on the steel balls.</p>
Full article ">Figure 7 Cont.
<p>The Raman spectrum of the coated samples. (<b>a</b>) The Raman spectra of the wear tracks. (<b>b</b>) The Raman spectra of the wear scars on the steel balls.</p>
Full article ">Figure 8
<p>The wear rate of (AlTiVCrNb)C<sub>x</sub>N<sub>y</sub> coating.</p>
Full article ">
12 pages, 10278 KiB  
Article
Enhanced Magnetocaloric Properties of the (MnNi)0.6Si0.62(FeCo)0.4Ge0.38 High-Entropy Alloy Obtained by Co Substitution
by Zhigang Zheng, Pengyan Huang, Xinglin Chen, Hongyu Wang, Shan Da, Gang Wang, Zhaoguo Qiu and Dechang Zeng
Entropy 2024, 26(9), 799; https://doi.org/10.3390/e26090799 - 19 Sep 2024
Viewed by 374
Abstract
In order to improve the magnetocaloric properties of MnNiSi-based alloys, a new type of high-entropy magnetocaloric alloy was constructed. In this work, Mn0.6Ni1−xSi0.62Fe0.4CoxGe0.38 (x = 0.4, 0.45, and 0.5) are [...] Read more.
In order to improve the magnetocaloric properties of MnNiSi-based alloys, a new type of high-entropy magnetocaloric alloy was constructed. In this work, Mn0.6Ni1−xSi0.62Fe0.4CoxGe0.38 (x = 0.4, 0.45, and 0.5) are found to exhibit magnetostructural first-order phase transitions from high-temperature Ni2In-type phases to low-temperature TiNiSi-type phases so that the alloys can achieve giant magnetocaloric effects. We investigate why chexagonal/ahexagonal (chexa/ahexa) gradually increases upon Co substitution, while phase transition temperature (Ttr) and isothermal magnetic entropy change (ΔSM) tend to gradually decrease. In particular, the x = 0.4 alloy with remarkable magnetocaloric properties is obtained by tuning Co/Ni, which shows a giant entropy change of 48.5 J∙kg−1K−1 at 309 K for 5 T and an adiabatic temperature change (ΔTad) of 8.6 K at 306.5 K. Moreover, the x = 0.55 HEA shows great hardness and compressive strength with values of 552 HV2 and 267 MPa, respectively, indicating that the mechanical properties undergo an effective enhancement. The large ΔSM and ΔTad may enable the MnNiSi-based HEAs to become a potential commercialized magnetocaloric material. Full article
(This article belongs to the Section Multidisciplinary Applications)
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the PPMS-based adiabatic temperature change direct measurement device.</p>
Full article ">Figure 2
<p>(<b>a</b>) The X-ray diffraction patterns of HEAs with different Co-doping at 295 K. (<b>b</b>) Unit cell parameters <span class="html-italic">c<sub>hex</sub>/a<sub>hex</sub></span> and volume <span class="html-italic">v</span> for Mn<sub>0.6</sub>Ni<sub>1−<span class="html-italic">x</span></sub>Si<sub>0.62</sub>Fe<sub>0.4</sub>Co<span class="html-italic"><sub>x</sub></span>Ge<sub>0.38</sub> (<span class="html-italic">x</span> = 0.4, 0.45, 0.5, 0.55) alloys determined from Rietveld refinements.</p>
Full article ">Figure 3
<p>DSC curves during heating process around <span class="html-italic">T<sub>C</sub></span> for HEAs with different Co-doping.</p>
Full article ">Figure 4
<p>Thermomagnetic curves of the HEAs with different Co-doping during the heating and cooling process at 0.05 T.</p>
Full article ">Figure 5
<p>Isothermal magnetization and demagnetization curves around <span class="html-italic">T<sub>C</sub></span> for (<b>a</b>) <span class="html-italic">x</span> = 0.4, (<b>b</b>) <span class="html-italic">x</span> = 0.45, (<b>c</b>) <span class="html-italic">x</span> = 0.5, (<b>d</b>) <span class="html-italic">x</span> = 0.55. The red arrows indicate magnetization, and the blue arrows indicate demagnetization.</p>
Full article ">Figure 6
<p>Three—dimensional surfaces showing −Δ<span class="html-italic">S<sub>M</sub></span> of (<b>a</b>) <span class="html-italic">x</span> = 0.4, (<b>b</b>) <span class="html-italic">x</span> = 0.45, (<b>c</b>) <span class="html-italic">x</span> = 0.5, (<b>d</b>) <span class="html-italic">x</span> = 0.55 under Δ<span class="html-italic">H</span> from 1 T to 5 T. The plots with the contour map in the plane of −Δ<span class="html-italic">S<sub>M</sub></span> are projected from 3D surfaces.</p>
Full article ">Figure 7
<p>The thermal hysteresis, −Δ<span class="html-italic">S<sub>M</sub></span> and <span class="html-italic">T<sub>C</sub></span> diagrams of HEAs with different Co-doping.</p>
Full article ">Figure 8
<p>Adiabatic temperature curves of <span class="html-italic">x</span> = 0.4 HEA and as a reference Gd under a 5 T magnetic field: (<b>a</b>) PPMS superconducting magnetic field; (<b>b</b>) 4.8 T pulsed magnetic field.</p>
Full article ">Figure 9
<p>Scatter-line plots of exponent <span class="html-italic">n</span> with respect to temperatures for Mn<sub>0.6</sub>Ni<sub>1−<span class="html-italic">x</span></sub>Si<sub>0.62</sub>Fe<sub>0.4</sub>Co<span class="html-italic"><sub>x</sub></span>Ge<sub>0.38</sub> (<span class="html-italic">x</span> = 0.4, 0.45, 0.5, 0.55) alloys.</p>
Full article ">Figure 10
<p>(<b>a</b>) Vickers hardness of HEAs with different Co-doping and as a reference gadolinium mental. (<b>b</b>) Compressive stress–strain curves of HEAs with different Co-doping.</p>
Full article ">
18 pages, 5385 KiB  
Article
High-Temperature Oxidation of NbTi-Bearing Refractory Medium- and High-Entropy Alloys
by Wei-Chih Lin, Yi-Wen Lien, Louis Etienne Moreau, Hideyuki Murakami, Kai-Chi Lo, Stéphane Gorsse and An-Chou Yeh
Materials 2024, 17(18), 4579; https://doi.org/10.3390/ma17184579 - 18 Sep 2024
Viewed by 569
Abstract
The oxidation of six NbTi-i refractory medium- and high-entropy alloys (NbTi + Ta, NbTi + CrTa, NbTi + AlTa, NbTi + AlMo, NbTi + AlMoTa and NbTi + AlCrMo) were investigated at 1000 °C for 20 h. According to our observation, increased Cr [...] Read more.
The oxidation of six NbTi-i refractory medium- and high-entropy alloys (NbTi + Ta, NbTi + CrTa, NbTi + AlTa, NbTi + AlMo, NbTi + AlMoTa and NbTi + AlCrMo) were investigated at 1000 °C for 20 h. According to our observation, increased Cr content promoted the formation of protective CrNbO4 and Cr2O3 oxides in NbTi + CrTa and NbTi + AlCrMo, enhancing oxidation resistance. The addition of Al resulted in the formation of AlTi-rich oxide in NbTi + AlTa. Ta addition resulted in the formation of complex oxides (MoTiTa8O25 and TiTaO4) and decreased oxidation resistance. Meanwhile, Mo’s low oxygen solubility could be beneficial for oxidation resistance while protective Cr2O3/CrNbO4 layers were formed. In NbTi + Ta, NbTi + AlTa and NbTi + CrTa, a considerable quantity of Ti-rich oxide was observed at the interdendritic region. In NbTi + AlCrMo, the enrichment of Cr and Ti at the interdendritic region could fasten the rate of oxidation. Compared to the recent research, NbTi + AlCrMo alloy is a light-weight oxidation-resistant alloy (weight gain of 1.29 mg/cm2 at 1000 °C for 20 h and low density (7.2 g/cm3)). Full article
Show Figures

Figure 1

Figure 1
<p>The statistics of element presence in RHEAs from the literature collection [<a href="#B7-materials-17-04579" class="html-bibr">7</a>,<a href="#B13-materials-17-04579" class="html-bibr">13</a>,<a href="#B14-materials-17-04579" class="html-bibr">14</a>,<a href="#B15-materials-17-04579" class="html-bibr">15</a>,<a href="#B16-materials-17-04579" class="html-bibr">16</a>,<a href="#B17-materials-17-04579" class="html-bibr">17</a>,<a href="#B19-materials-17-04579" class="html-bibr">19</a>,<a href="#B20-materials-17-04579" class="html-bibr">20</a>,<a href="#B21-materials-17-04579" class="html-bibr">21</a>,<a href="#B22-materials-17-04579" class="html-bibr">22</a>,<a href="#B25-materials-17-04579" class="html-bibr">25</a>,<a href="#B27-materials-17-04579" class="html-bibr">27</a>,<a href="#B28-materials-17-04579" class="html-bibr">28</a>,<a href="#B29-materials-17-04579" class="html-bibr">29</a>,<a href="#B30-materials-17-04579" class="html-bibr">30</a>,<a href="#B31-materials-17-04579" class="html-bibr">31</a>,<a href="#B32-materials-17-04579" class="html-bibr">32</a>,<a href="#B33-materials-17-04579" class="html-bibr">33</a>,<a href="#B34-materials-17-04579" class="html-bibr">34</a>,<a href="#B35-materials-17-04579" class="html-bibr">35</a>,<a href="#B36-materials-17-04579" class="html-bibr">36</a>,<a href="#B37-materials-17-04579" class="html-bibr">37</a>,<a href="#B38-materials-17-04579" class="html-bibr">38</a>,<a href="#B39-materials-17-04579" class="html-bibr">39</a>,<a href="#B40-materials-17-04579" class="html-bibr">40</a>,<a href="#B41-materials-17-04579" class="html-bibr">41</a>,<a href="#B42-materials-17-04579" class="html-bibr">42</a>,<a href="#B43-materials-17-04579" class="html-bibr">43</a>,<a href="#B44-materials-17-04579" class="html-bibr">44</a>,<a href="#B45-materials-17-04579" class="html-bibr">45</a>,<a href="#B46-materials-17-04579" class="html-bibr">46</a>,<a href="#B47-materials-17-04579" class="html-bibr">47</a>,<a href="#B48-materials-17-04579" class="html-bibr">48</a>,<a href="#B49-materials-17-04579" class="html-bibr">49</a>,<a href="#B50-materials-17-04579" class="html-bibr">50</a>,<a href="#B51-materials-17-04579" class="html-bibr">51</a>,<a href="#B52-materials-17-04579" class="html-bibr">52</a>,<a href="#B53-materials-17-04579" class="html-bibr">53</a>,<a href="#B54-materials-17-04579" class="html-bibr">54</a>,<a href="#B55-materials-17-04579" class="html-bibr">55</a>].</p>
Full article ">Figure 2
<p>The XRD patterns of NbTi-bearing RMEAs and RHEAs.</p>
Full article ">Figure 3
<p>The as-cast microstructure of (<b>a</b>) NbTi + Ta, (<b>b</b>) NbTi + CrTa, (<b>c</b>) NbTi + AlMo, (<b>d</b>) NbTi + AlTa, (<b>e</b>) NbTi + AlCrMo and (<b>f</b>) NbTi + AlMoTa near the surface area.</p>
Full article ">Figure 4
<p>The oxidation mass gain curve tested at 1000 °C for 20 h.</p>
Full article ">Figure 5
<p>The oxidized NbTi + AlCrMo for 20 h at 1000 °C: (<b>a</b>) a general view of the oxide layer with EDS mapping, the magnified SEM image of (<b>b</b>) the external oxidation region and (<b>c</b>) the internal oxidation region and (<b>d</b>) the XRD pattern of the oxide layer. (<b>e</b>) The EBSD image with the band contrast and phase characterization of the external oxide region. The red arrows in (<b>a</b>) indicates the oxidation of the interdendritic region. The compositions of the indicated regions in (<b>b</b>,<b>c</b>) are listed in <a href="#materials-17-04579-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 6
<p>The XRD patterns of NbTi + Ta, NbTi + CrTa and NbTi + AlTa after oxidation at 1000 °C for 20 h.</p>
Full article ">Figure 7
<p>The BEI images after oxidation at 1000 °C for 20 h: a general view of the oxide layer of (<b>a</b>) NbTi + Ta, (<b>d</b>) NbTi + CrTa and (<b>g</b>) NbTi + AlTa, the magnified SEM image of the external oxidation region of (<b>b</b>) NbTi + Ta, (<b>e</b>) NbTi + CrTa and (<b>h</b>) NbTi + AlTa and the internal oxidation region of (<b>c</b>) NbTi + Ta, (<b>f</b>) NbTi + CrTa and (<b>i</b>) NbTi + AlTa. The compositions of the indicated regions in (<b>b</b>,<b>c</b>,<b>e</b>–<b>h</b>) are listed in <a href="#materials-17-04579-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 8
<p>The EBSD band contrast and phase characterization images of (<b>a</b>) NbTi + Ta, (<b>b</b>) NbTi + CrTa and (<b>c</b>) NbTi + AlTa.</p>
Full article ">Figure 9
<p>The XRD patterns of NbTi + AlMo and NbTi + AlMoTa after oxidation at 1000 °C for 20 h.</p>
Full article ">Figure 10
<p>(<b>a</b>) The oxide layer BEI with EDS mapping and (<b>b</b>) the magnified image of NbTi + AlMo after 1000 °C oxidation for 20 h. (<b>c</b>) The oxide layer BEI of NbTi + AlMoTa after 1000 °C oxidation for 20 h. The BEI of the interface between the oxide layer and oxygen-affected zone for (<b>d</b>) NbTi + AlMo and (<b>e</b>) NbTi + AlMoTa after 1000 °C oxidation for 20 h. The compositions of the indicated regions in (<b>b</b>,<b>c</b>) are listed in <a href="#materials-17-04579-t007" class="html-table">Table 7</a>.</p>
Full article ">Figure 11
<p>The EBSD band contrast and phase characterization images of (<b>a</b>) NbTi + AlMo and (<b>b</b>) NbTi + AlMoTa.</p>
Full article ">Figure 12
<p>The comparison of the specific mass gain tested at 1000 °C for 20 h among NbTi-bearing RMEAs and RHEAs from this work, AlCrMo + Nb/NbTi/TaTi [<a href="#B25-materials-17-04579" class="html-bibr">25</a>], CrNbTiZr [<a href="#B28-materials-17-04579" class="html-bibr">28</a>] and AlHfMoNbTi [<a href="#B38-materials-17-04579" class="html-bibr">38</a>] from the literature.</p>
Full article ">
23 pages, 7411 KiB  
Review
Improvement of High Temperature Wear Resistance of Laser-Cladding High-Entropy Alloy Coatings: A Review
by Yantao Han and Hanguang Fu
Metals 2024, 14(9), 1065; https://doi.org/10.3390/met14091065 - 18 Sep 2024
Viewed by 972
Abstract
As a novel type of metal material emerging in recent years, high-entropy alloy boasts properties such as a simplified microstructure, high strength, high hardness and wear resistance. High-entropy alloys can use laser cladding to produce coatings that exhibit excellent metallurgical bonding with the [...] Read more.
As a novel type of metal material emerging in recent years, high-entropy alloy boasts properties such as a simplified microstructure, high strength, high hardness and wear resistance. High-entropy alloys can use laser cladding to produce coatings that exhibit excellent metallurgical bonding with the substrate, thereby significantly improvement of the wear resistance of the material surface. In this paper, the research progress on improving the high-temperature wear resistance of high entropy alloy coatings (LC-HEACs) was mainly analyzed based on the effect of some added alloying elements and the presence of hard ceramic phases. Building on this foundation, the study primarily examines the impact of adding elements such as aluminum, titanium, copper, silicon, and molybdenum, along with hard ceramic particles like TiC, WC, and NbC, on the phase structure of coatings, high-temperature mechanisms, and the synergistic interactions between these elements. Additionally, it explores the potential of promising lubricating particles and introduces an innovative, highly efficient additive manufacturing technology known as extreme high-speed laser metal deposition (EHLMD). Finally, this paper summarizes the main difficulties involved in increasing the high-temperature wear resistance of LC-HEACs and some problems worthy of attention in the future development. Full article
(This article belongs to the Special Issue Surface Engineering and Coating Tribology—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the research content.</p>
Full article ">Figure 2
<p>SEM images of the FeCoCrNiMnAl<sub>x</sub> coatings after high-temperature oxidation test: (<b>a</b>,<b>a1</b>,<b>a2</b>) x = 0; (<b>b</b>,<b>b1</b>,<b>b2</b>) x = 0.25; (<b>c</b>,<b>c1</b>,<b>c2</b>) x = 0.5; (<b>d</b>,<b>d1</b>,<b>d2</b>) x = 0.75. (Reprinted with permission from [<a href="#B20-metals-14-01065" class="html-bibr">20</a>]. 2020 Elsevier).</p>
Full article ">Figure 3
<p>Standard Gibbs free energy vs. temperature curves for reactions (3)-(8). (Reprinted with permission from [<a href="#B20-metals-14-01065" class="html-bibr">20</a>]. 2020 Elsevier.)</p>
Full article ">Figure 4
<p>Load–displacement curves of the FeCoCrNiMnAlx cladding layers after the nanoindentation test. (Reprinted with permission from [<a href="#B26-metals-14-01065" class="html-bibr">26</a>]. 2019 Elsevier).</p>
Full article ">Figure 5
<p>XRD patterns of AlCoCrFeNi after wear testing at 300 °C, 600 °C, 800 °C and 900 °C. (Reprinted with permission from [<a href="#B31-metals-14-01065" class="html-bibr">31</a>]. 2023 Elsevier).</p>
Full article ">Figure 6
<p>Cross-sectional micromorphologies and EDS elemental analysis of the HEA coatings and 45 # steel after high-temperature oxidation for 20 h: (<b>a</b>) x = 0, (<b>b</b>) x = 0.25, (<b>c</b>) x = 0.5, (<b>d</b>) x = 0.75, (<b>e</b>) x = 1.0 and (<b>f</b>) 45# steel. (Reprinted with permission from [<a href="#B45-metals-14-01065" class="html-bibr">45</a>]. 2018 Elsevier).</p>
Full article ">Figure 7
<p>(<b>a</b>) Microstructure of CoCrFeNiCu alloy after wear test at 600 °C. (<b>b</b>) Microhardness and wear rate as a function of Cu content. (Reprinted with permission from [<a href="#B49-metals-14-01065" class="html-bibr">49</a>]. 2023 Elsevier).</p>
Full article ">Figure 8
<p>Diagram of wear process of FeCoCrNiCu high-entropy alloy coating at 600 °C. (Reprinted with permission from [<a href="#B50-metals-14-01065" class="html-bibr">50</a>]. 2024 Elsevier).</p>
Full article ">Figure 9
<p>The dark-field TEM images (<b>a</b>,<b>b</b>) and the bright-field TEM images (<b>c</b>) of the Mo<sub>1.0</sub>NbTiZr RHEA coating; HRTEM images (<b>d</b>–<b>f</b>); SAED patterns (<b>g</b>–<b>j</b>) of the different phases. (<b>a</b>–<b>c</b>,<b>f</b>) correspond to different regions of the Mo1.0 coating, and (<b>d</b>,<b>e</b>) are magnified views of (<b>a</b>,<b>b</b>), respectively. (Reprinted with permission from [<a href="#B56-metals-14-01065" class="html-bibr">56</a>]. 2023 Elsevier).</p>
Full article ">Figure 10
<p>T4 coating characterized by TEM (<b>a</b>) morphology; (<b>b</b>) elongated TiC; (<b>c</b>) TiC of different shapes; (<b>d</b>) element statistics histogram of Spot 1, Spot 2 and Spot 3. (Reprinted with permission from [<a href="#B63-metals-14-01065" class="html-bibr">63</a>]. 2022 Elsevier).</p>
Full article ">Figure 11
<p>Microstructure of composite coatings with different TiC content and the intermediate layer formed during the high-temperature wear: (<b>a</b>,<b>b</b>) T5C and T10C; (<b>c</b>,<b>d</b>) T15C and T20C. (Reprinted with permission from [<a href="#B65-metals-14-01065" class="html-bibr">65</a>]. 2022 Elsevier).</p>
Full article ">Figure 12
<p>SEM images of the worn surface of x wt.% WC/AlCoCrFeNiTi<sub>0.5</sub> HEAs at the applied load of 5 N at HT. (<b>a</b>) 0%; (<b>b</b>) 5%; (<b>c</b>) 8%; (<b>d</b>) 10%; (<b>e</b>) 15%. (Reprinted with permission from [<a href="#B71-metals-14-01065" class="html-bibr">71</a>]. 2024 Elsevier).</p>
Full article ">Figure 13
<p>Wear rate of the RHEAs coatings at 25, 600, and 800 °C. (Reprinted with permission from [<a href="#B74-metals-14-01065" class="html-bibr">74</a>]. 2023 Elsevier).</p>
Full article ">Figure 14
<p>Schematic diagram of structural evolution of (CuNiTiNbCr)C<sub>x</sub> HEFs with carbon content. (Reprinted with permission from [<a href="#B77-metals-14-01065" class="html-bibr">77</a>]. 2024 Elsevier).</p>
Full article ">Figure 15
<p>(<b>a</b>) Friction coefficient curves, (<b>b</b>) average friction coefficients, (<b>c</b>) wear track profiles, and (<b>d</b>) wear rates of (CuNiTiNbCr)C<sub>x</sub> HEFs with different carbon content. (Reprinted with permission from [<a href="#B77-metals-14-01065" class="html-bibr">77</a>]. 2024 Elsevier).</p>
Full article ">Figure 16
<p>Wear volumes (<b>a</b>) and wear rates (<b>b</b>) of CLMD and EHLMD K648 superalloy under different loads. (Reprinted with permission from [<a href="#B83-metals-14-01065" class="html-bibr">83</a>]. 2024 Elsevier).</p>
Full article ">
11 pages, 3307 KiB  
Article
Influence of Oxygen and Nitrogen Flow Ratios on the Microstructure Evolution in AlCrTaTiZr High-Entropy Oxynitride Films
by Yung-Chu Liang, Ching-Yin Lee, Miao-I Lin, Ting-En Shen, Jung-Fan Hung, Jien-Wei Yeh and Che-Wei Tsai
Coatings 2024, 14(9), 1199; https://doi.org/10.3390/coatings14091199 - 18 Sep 2024
Viewed by 355
Abstract
This study explores the influence of oxygen and nitrogen flow ratios on the microstructure and mechanical properties of AlCrTaTiZr high-entropy oxynitride films. Oxygen flow rates (0%–0.75%) were adjusted while maintaining a fixed nitrogen flow ratio (RN = 15%) to fabricate films with [...] Read more.
This study explores the influence of oxygen and nitrogen flow ratios on the microstructure and mechanical properties of AlCrTaTiZr high-entropy oxynitride films. Oxygen flow rates (0%–0.75%) were adjusted while maintaining a fixed nitrogen flow ratio (RN = 15%) to fabricate films with similar compositions. The results show that increasing oxygen flow enhanced hardness through solid solution strengthening and grain refinement, though excessive oxygen caused an amorphous structure and reduced hardness. After annealing at 900 °C, the hardness of all films was further increased. The film with a nitrogen flow ratio 40 times higher than oxygen exhibited the highest hardness of 21.8 GPa, along with superior mechanical performance. These findings highlight the potential of high-entropy oxynitride films for applications requiring high wear resistance and adhesion. Full article
(This article belongs to the Special Issue High Entropy Alloy Films)
Show Figures

Figure 1

Figure 1
<p>Chemical composition of the films as a function of the oxygen flow ratio when R<sub>N</sub> is fixed at 15%.</p>
Full article ">Figure 2
<p>GIXRD patterns of films at various nitrogen flow ratios, as the R<sub>N</sub> is fixed at 15%: (<b>a</b>) as-deposited; (<b>b</b>) annealed at 900 °C for 5 h.</p>
Full article ">Figure 3
<p>XPS spectra of as-deposited specimens are prepared at R<sub>O</sub> = 0.75%, and the bonding fraction of each metal reacts with reactive gasses: (<b>a</b>) Al, (<b>b</b>) Cr, (<b>c</b>) Ta, (<b>d</b>) Ti, and (<b>e</b>) Zr.</p>
Full article ">Figure 4
<p>The top surface and cross-section morphology of the as-deposited films at different oxygen flow ratios as R<sub>N</sub> is fixed 15%; (<b>a</b>,<b>f</b>) sample L0; (<b>b</b>,<b>g</b>) sample L1; (<b>c</b>,<b>h</b>) sample L2; (<b>d</b>,<b>i</b>) sample L3; (<b>e</b>,<b>j</b>) sample L4.</p>
Full article ">Figure 5
<p>Hardness and modulus of as-deposited and annealed films at different oxygen flow ratios.</p>
Full article ">Figure 6
<p>(<b>a</b>) H/E and (<b>b</b>) H<sup>3</sup>/E<sup>2</sup> ratio of the as-deposited and annealed films as a function of oxygen flow ratio when R<sub>N</sub> is fixed at 15%.</p>
Full article ">
10 pages, 2750 KiB  
Article
Carbon Nanofiber-Encapsulated FeCoNiCuMn Sulfides with Tunable S Doping for Enhanced Oxygen Evolution Reaction
by Yuhan Sun, Chen Shen, Mingran Wang, Yang Cao, Qianwei Wang, Jiayi Rong, Tong He, Duanyang Li and Feng Cao
Catalysts 2024, 14(9), 626; https://doi.org/10.3390/catal14090626 - 17 Sep 2024
Viewed by 417
Abstract
The oxygen evolution reaction (OER) stands out as a key electrochemical process for the conversion of clean energy. However, the practical implementation of OER is frequently impeded by its slow kinetics and the necessity for scarce and expensive noble metal catalysts. High-entropy transition [...] Read more.
The oxygen evolution reaction (OER) stands out as a key electrochemical process for the conversion of clean energy. However, the practical implementation of OER is frequently impeded by its slow kinetics and the necessity for scarce and expensive noble metal catalysts. High-entropy transition metal sulfides (HETMS) stand at the forefront of OER catalysts, renowned for their exceptional catalytic performance and diversity. Herein, we have synthesized a HETMS catalyst, (FeCoNiCuMn50)S2, encapsulated within carbon nanofibers through a one-step process involving the synergistic application of electrospinning and chemical vapor deposition. By precisely controlling the doping levels of sulfur, we have demonstrated that sulfur incorporation significantly increases the exposed surface area of alloy particles on carbon nanofibers and optimizes the electronic configuration of the alloy elements. These findings reveal that sulfur doping is instrumental in the substantial improvement of the catalyst’s OER performance. Notably, the catalyst showed optimal activity at a sulfur-to-metal atom ratio of 2:1, delivering an overpotential of 254 mV at a current density of 10 mA cm−2 in 1.0 M KOH solution. Furthermore, the (FeCoNiCuMn50)S2 catalyst exhibited remarkable electrochemical stability, underscoring its potential as an efficient and robust OER electrocatalyst for sustainable energy applications. Full article
(This article belongs to the Section Catalytic Materials)
Show Figures

Figure 1

Figure 1
<p>Characterization of high-entropy metal sulfide (FeCoNiCuMn<sub>50</sub>)S<sub>2</sub> nanoparticles. (<b>a</b>) Schematic of the synthesis process; (<b>b</b>) SEM image; (<b>c</b>) TEM image and particle size distribution (inset); (<b>d</b>) HRTEM image and SAED image (inset); (<b>e</b>) STEM and corresponding mapping images; and (<b>f</b>) XRD of (FeCoNiCuMn<sub>50</sub>)S<sub>x</sub>.</p>
Full article ">Figure 2
<p>XPS spectra of (<b>a</b>) Fe 2p; (<b>b</b>) Co 2p; (<b>c</b>) Ni 2p; (<b>d</b>) Cu 2p; (<b>e</b>) Mn 2p; and (<b>f</b>) S 2p of np-HETMS (FeCoNiCuMn<sub>50</sub>)S<sub>2</sub> nanoparticles.</p>
Full article ">Figure 3
<p>Electrocatalytic performance of different samples. (<b>a</b>) OER polarization curves of different samples. (<b>b</b>) Tafel plots. (<b>c</b>) Nyquist plots for different samples. (<b>d</b>) Capacitive currents as a function of scan rate. (<b>e</b>) Performance diagram of OER electrocatalysts. (<b>f</b>) Stability testing.</p>
Full article ">
17 pages, 6501 KiB  
Article
Enhancing Mechanical Properties of Graphene/Aluminum Nanocomposites via Microstructure Design Using Molecular Dynamics Simulations
by Zhonglei Ma, Hongding Wang, Yanlong Zhao, Zhengning Li, Hong Liu, Yizhao Yang and Zigeng Zhao
Materials 2024, 17(18), 4552; https://doi.org/10.3390/ma17184552 - 16 Sep 2024
Viewed by 572
Abstract
This study explores the mechanical properties of graphene/aluminum (Gr/Al) nanocomposites through nanoindentation testing performed via molecular dynamics simulations in a large-scale atomic/molecular massively parallel simulator (LAMMPS). The simulation model was initially subjected to energy minimization at 300 K, followed by relaxation for 50 [...] Read more.
This study explores the mechanical properties of graphene/aluminum (Gr/Al) nanocomposites through nanoindentation testing performed via molecular dynamics simulations in a large-scale atomic/molecular massively parallel simulator (LAMMPS). The simulation model was initially subjected to energy minimization at 300 K, followed by relaxation for 50 ps under the NPT ensemble, wherein the number of atoms (N), simulation temperature (T), and pressure (P) were conserved. After the model was fully relaxed, loading and unloading simulations were performed. This study focused on the effects of the Gr arrangement with a brick-and-mortar structure and incorporation of high-entropy alloy (HEA) coatings on mechanical properties. The findings revealed that Gr sheets (GSs) significantly impeded dislocation propagation, preventing the dislocation network from penetrating the Gr layer within the plastic zone. However, interactions between dislocations and GSs in the Gr/Al nanocomposites resulted in reduced hardness compared with that of pure aluminum. After modifying the arrangement of GSs and introducing HEA (FeNiCrCoAl) coatings, the elastic modulus and hardness of the Gr/Al nanocomposites were 83 and 9.5 GPa, respectively, representing increases of 21.5% and 17.3% compared with those of pure aluminum. This study demonstrates that vertically oriented GSs in combination with HEA coatings at a mass fraction of 3.4% significantly enhance the mechanical properties of the Gr/Al nanocomposites. Full article
(This article belongs to the Section Materials Simulation and Design)
Show Figures

Figure 1

Figure 1
<p>Schematic of the nanoindentation models: pure aluminum (<b>a</b>), Gr/Al-level3 (<b>b</b>), Gr/Al-vertical3 (<b>c</b>), HEA/Gr/Al-level3 (<b>d</b>), and HEA/Gr/Al-vertical3 (<b>e</b>).</p>
Full article ">Figure 2
<p>Dislocation distributions observed along the x-axis for the five models, namely pure aluminum, Gr/Al-level3, Gr/Al-vertical3, HEA/Gr/Al-level3, and HEA/Gr/Al-vertical3, at indenter displacements of d = 15, 20, 25, 30, and 35 Å. Dislocations are colored according to their Burgers vector. Green: 1/6&lt;112&gt;; Dark blue: 1/2&lt;110&gt;; Pink: 1/6&lt;110&gt;; Yellow: 1/3&lt;100&gt;; Bright blue: 1/3&lt;111&gt;; Red: others.</p>
Full article ">Figure 3
<p>Side views of the y–z-planes of Gr/Al-level3 (<b>a</b>) and HEA/Gr/Al-level3 (<b>b</b>) at indentation depths of 0, 20, 25, and 30 Å. The viewing direction is along the x-axis.</p>
Full article ">Figure 4
<p>Side views of the y–z-planes of Gr/Al-vertical3 (<b>a</b>) at indentation depths of 0, 26, 27, 28, and 29 Å and HEA/Gr/Al-vertical3 (<b>b</b>) at indentation depths of 0, 9, 11, 13, and 16 Å. The viewing direction is along the x-axis.</p>
Full article ">Figure 5
<p>In-plane height profiles of Gr at an indentation depth of 30 Å for models Gr/Al-level3 (<b>a</b>), HEA/Gr/Al-level3 (<b>b</b>), Gr/Al-vertical3 (<b>c</b>), and HEA/Gr/Al-vertical3 (<b>d</b>).</p>
Full article ">Figure 6
<p>Evolution of total dislocation length (<b>a</b>) and indentation force (<b>b</b>) with indenter displacement.</p>
Full article ">Figure 7
<p>Hardness values vs. indentation depths.</p>
Full article ">Figure 8
<p>Force–displacement curves at the unloading stage.</p>
Full article ">Figure 9
<p>Distribution of dislocation lines and defect atoms at different indenter displacements during the unloading stage. Dislocations are colored according to their Burgers vector. Green: 1/6&lt;112&gt;; Dark blue: 1/2&lt;110&gt;; Pink: 1/6&lt;110&gt;; Yellow: 1/3&lt;100&gt;; Bright blue: 1/3&lt;111&gt;; Red: others. (<b>a</b>) pure aluminum; (<b>b</b>) Gr/Al-level3; (<b>c</b>) Gr/Al-vertical3; (<b>d</b>) HEA/Gr/Al-level3; (<b>e</b>) HEA/Gr/Al-vertical3.</p>
Full article ">Figure 10
<p>Reduced Young’s modulus of the five models.</p>
Full article ">Figure 11
<p>Dislocation length–indenter displacement curves. (<b>a</b>) pure aluminum; (<b>b</b>) Gr/Al-vertical3; (<b>c</b>) HEA/Gr/Al-level3; (<b>d</b>) HEA/Gr/Al-vertical3.</p>
Full article ">
14 pages, 9055 KiB  
Article
Ultra-High Strength in FCC+BCC High-Entropy Alloy via Different Gradual Morphology
by Ziheng Ding, Chaogang Ding, Zhiqin Yang, Hao Zhang, Fanghui Wang, Hushan Li, Jie Xu, Debin Shan and Bin Guo
Materials 2024, 17(18), 4535; https://doi.org/10.3390/ma17184535 - 15 Sep 2024
Viewed by 419
Abstract
In this study, high-pressure torsion (HPT) processing is applied to the as-cast Al0.5CoCrFeNi high-entropy alloy (HEA) for 1, 3, and 5 turns. Microstructural observations reveal a significant refinement of the second phase after HPT processing. This refinement effect is influenced by [...] Read more.
In this study, high-pressure torsion (HPT) processing is applied to the as-cast Al0.5CoCrFeNi high-entropy alloy (HEA) for 1, 3, and 5 turns. Microstructural observations reveal a significant refinement of the second phase after HPT processing. This refinement effect is influenced by the number of processing turns and the distance of the processing position from the center. As the number of processing turns or the distance of the processing position from the center increases, the fragmentation effect on the second phase becomes more pronounced. The hardness of the alloy is greatly enhanced after HPT processing, but there is an upper limit to this enhancement. After increasing the number of processing turns to 5, the increase in hardness at the edge becomes less significant, while the overall hardness becomes more uniform. Additionally, the strength of the processed alloy is significantly enhanced, while its ductility undergoes a noticeable decrease. With an increase in the number of processing turns, the second phase is further refined, resulting in improvement of strength and ductility. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic illustration of (<b>a</b>) HPT method and (<b>b</b>) sample and procedure used for different characterization methods.</p>
Full article ">Figure 2
<p>(<b>a</b>) XRD analysis for the as-cast sample and after HPT through 1, 3, and 5 turns; (<b>b</b>) grain size and second-phase size under different processing cycles.</p>
Full article ">Figure 3
<p>Microstructure and element distribution map of Al<sub>0.5</sub>CoCrFeNi as-cast alloy.</p>
Full article ">Figure 4
<p>SEM-BSE images showing the microstructure at the center, 1/2 of radius, and edge region after HPT for 1, 3, and 5 turns.</p>
Full article ">Figure 5
<p>Element distribution map at the center after HPT for 1 turn.</p>
Full article ">Figure 6
<p>The Vickers microhardness plotted against (<b>a</b>) 0.5 turns, (<b>b</b>) 1 turn, (<b>c</b>) 3 turns, (<b>d</b>) distance from the disk center, and (<b>e</b>) equivalent strain.</p>
Full article ">Figure 7
<p>Tensile properties of Al<sub>0.5</sub>CoCrFeNi HEAs processed by HPT: (<b>a</b>) engineering stress–strain curves; (<b>b</b>) changing tendencies of yield strength, tensile strength, and elongation; (<b>c</b>) work hardening rate plotted against true strain; (<b>d</b>) comparison with the data collected from other studies [<a href="#B58-materials-17-04535" class="html-bibr">58</a>,<a href="#B59-materials-17-04535" class="html-bibr">59</a>,<a href="#B60-materials-17-04535" class="html-bibr">60</a>].</p>
Full article ">Figure 8
<p>The microscopic morphology of fracture surface: (<b>a</b>) as-cast; (<b>b</b>) 1 turn; (<b>c</b>) 3 turns; (<b>d</b>) 5 turns.</p>
Full article ">
11 pages, 11934 KiB  
Article
Effect of Alloying on Microstructure and Mechanical Properties of AlCoCrFeNi2.1 Eutectic High-Entropy Alloy
by Xue-Yao Tian, Hong-Liang Zhang, Zhi-Sheng Nong, Xue Cui, Ze-Hao Gu, Teng Liu, Hong-Mei Li and Eshkuvat Arzikulov
Materials 2024, 17(18), 4471; https://doi.org/10.3390/ma17184471 - 12 Sep 2024
Viewed by 418
Abstract
In order to explore the effect of alloying on the microstructures and mechanical properties of AlCoCrFeNi2.1 eutectic high-entropy alloys (EHEAs), 0.1, 0.2, and 0.3 at.% V, Mo, and B were added to the AlCoCrFeNi2.1 alloy in this work. The effects of [...] Read more.
In order to explore the effect of alloying on the microstructures and mechanical properties of AlCoCrFeNi2.1 eutectic high-entropy alloys (EHEAs), 0.1, 0.2, and 0.3 at.% V, Mo, and B were added to the AlCoCrFeNi2.1 alloy in this work. The effects of the elements and contents on the phase composition, microstructures, mechanical properties, and fracture mechanism were investigated. The results showed that the crystal structures of the AlCoCrFeNi2.1 EHEAs remained unchanged, and the alloys were still composed of FCC and BCC structures, whose content varied with the addition of alloying elements. After alloying, the aggregation of Co, Cr, Al, and Ni elements remained unchanged, and the V and Mo were distributed in both dendritic and interdendritic phases. The tensile strengths of the alloys all exceeded 1000 MPa when the V and Mo elements were added, and the Mo0.2 alloy had the highest tensile strength, of 1346.3 MPa, and fracture elongation, of 24.6%. The alloys with the addition of V and Mo elements showed a mixed ductile and brittle fracture, while the B-containing alloy presented a cleavage fracture. The fracture mechanism of Mo0.2 alloy is mainly crack propagation in the BCC lamellae, and the FCC dendritic lamellae exhibit the characteristics of plastic deformation. Full article
(This article belongs to the Special Issue Advanced Science and Technology of High Entropy Materials)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD diffractions patterns (<b>a</b>) and microstructures (<b>b</b>) of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) EHEAs.</p>
Full article ">Figure 2
<p>SEM images of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) EHEAs (<b>a</b>), and element distribution of V0.2 and Mo0.2 alloys (<b>b</b>).</p>
Full article ">Figure 2 Cont.
<p>SEM images of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) EHEAs (<b>a</b>), and element distribution of V0.2 and Mo0.2 alloys (<b>b</b>).</p>
Full article ">Figure 3
<p>Hardness profile corresponding to the depth of penetration and hardness of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) alloys.</p>
Full article ">Figure 4
<p>Stress–strain curves and fracture topography of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) alloys (<b>a</b>) and fracture mechanisms of V0.2 and Mo0.2 alloys (<b>b</b>).</p>
Full article ">Figure 4 Cont.
<p>Stress–strain curves and fracture topography of AlCoCrFeNi<sub>2.1</sub>X<sub>y</sub> (X = V, Mo, B; y = 0.1, 0.2, 0.3) alloys (<b>a</b>) and fracture mechanisms of V0.2 and Mo0.2 alloys (<b>b</b>).</p>
Full article ">Figure 5
<p>Distribution of grain orientation and phase composition of Mo0.2 and B0.1 alloys after tensile test.</p>
Full article ">
13 pages, 5969 KiB  
Article
Abnormal Effect of Al on the Phase Stability and Deformation Mechanism of Ti-Zr-Hf-Al Medium-Entropy Alloys
by Penghao Yuan, Lu Wang, Ying Liu and Xidong Hui
Metals 2024, 14(9), 1035; https://doi.org/10.3390/met14091035 - 11 Sep 2024
Viewed by 420
Abstract
Complex concentrated alloys, including high-entropy alloys (HEAs) and medium-entropy alloys (MEAs), offer another pathway for developing metals with excellent mechanical properties. However, HEAs/MEAs of different structures often suffer from various drawbacks. So, investigations on the effect of phase and microstructure on their properties [...] Read more.
Complex concentrated alloys, including high-entropy alloys (HEAs) and medium-entropy alloys (MEAs), offer another pathway for developing metals with excellent mechanical properties. However, HEAs/MEAs of different structures often suffer from various drawbacks. So, investigations on the effect of phase and microstructure on their properties become necessary. In the present work, we adjust the phase constitution and microstructure by Al addition in a series of (Ti2ZrHf)100−xAlx (x = 12, 14, 16, 18, 20, at.%, named Alx) MEAs. Different from traditional titanium, Al shows a β-stabilizing effect, and the phase follows the evolution of α′(α)→α″→β + ω + B2 with Al increasing from 12 to 20 at.%, which could not be predicted by the CALPHAD (Calculate Phase Diagrams) method or the Bo-Md diagram because of the complex interactions among composition elements. At a low Al content, the solid solution strengthening of the HCP phase contributes to the extremely high strength with a σ0.2 of 1528 MPa and σb of 1937 MPa for Al14. The appearance of α″ deteriorates the deformation capability with increasing Al content in the Al16 and Al18 MEAs. In the Al20 MEA, Al improves the formations of ordered B2 and metastable β. The phase transformation strengthening, including B2 to BCC and BCC to α″, together with the precipitation strengthening of ω, brings about a high work-hardening ratio (above 5 GPa) and improvements in ductility (6.8% elongation). This work provides guidelines for optimizing the properties of MEAs. Full article
(This article belongs to the Section Entropic Alloys and Meta-Metals)
Show Figures

Figure 1

Figure 1
<p>The phase content predicted by Thermo-calc. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20. The legend in (<b>e</b>) applies to (<b>a</b>–<b>d</b>).</p>
Full article ">Figure 2
<p>The phase prediction by the Bo-Md method.</p>
Full article ">Figure 3
<p>The XRD diffractions of Ti-Zr-Hf-Al MEAs.</p>
Full article ">Figure 4
<p>The microstructure of Ti-Zr-Hf-Al MEAs. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20.</p>
Full article ">Figure 5
<p>The microstructure investigated by TEM for the Al12 MEA. (<b>a</b>) The HAADF image and EDS elements mapping; (<b>b</b>) typical bright-field TEM images showing the laths with different sizes in the Al12 alloy, together with the inset image of SAED for the region marked by the red circle; (<b>c</b>) the HRTEM image of the area marked by the yellow square in (<b>c</b>); (<b>d</b>) the FFT of (<b>c</b>); the FFT (<b>e1</b>) and amplified image (<b>e2</b>) for the area e in (<b>c</b>); and the FFT (<b>f1</b>) and amplified image (<b>f2</b>) for the area f in (<b>c</b>).</p>
Full article ">Figure 6
<p>The microstructure investigated by TEM for the Al20 MEA. (<b>a</b>) The HAADF image and EDS element mapping; (<b>b1</b>) typical bright-field TEM image for the Al20 alloy; (<b>b2</b>) SAED for the region marked by the red circle in (<b>b1</b>); (<b>b3</b>) the dark-field TEM image of the ω phase shown in (<b>b2</b>), (<b>b4</b>) the HRTEM image of the α″ and the BCC matrix; (<b>c1</b>) the HRTEM image of the ω phase and (<b>c2</b>) the responding FFT; and (<b>c3</b>) the HRTEM image of the B2 phase and (<b>c4</b>) the responding FFT.</p>
Full article ">Figure 7
<p>The mechanical properties of Al<sub>x</sub> MEAs at room temperature. (<b>a</b>) Tensile curves at room temperature; (<b>b</b>) the compression curves of Al20; and the corresponding work-hardening rate curves of (<b>c</b>) Al12 and (<b>d</b>) Al20.</p>
Full article ">Figure 8
<p>The fracture morphologies of the Al<sub>x</sub> MEAs. (<b>a</b>) Al12; (<b>b</b>) Al14; (<b>c</b>) Al16; (<b>d</b>) Al18; and (<b>e</b>) Al20.</p>
Full article ">Figure 9
<p>The DSC curves of the Al12 and Al20 MEAs.</p>
Full article ">Figure 10
<p>The XRD patterns of Al20 before and after the tensile test.</p>
Full article ">
Back to TopTop